The Pre - and Post-Failure Deformation Behavior of Soil Slopes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 1410

______________________________________________________________________

The Pre- and Post-Failure

Deformation Behaviour of Soil Slopes


______________________________________________________________________

Gavan James Hunter

A thesis submitted in partial fulfilment


of the requirements for the degree of
Doctor of Philosophy

School of Civil and Environmental Engineering

The University of New South Wales

April 2003
Abstract Page i

Abstract
This thesis examines the pre and post failure deformation behaviour of landslides in cut,
fill and natural soil slopes, and of the deformation behaviour of embankment dams. The
deformation behaviour of landslides and embankment dams have been analysed from a
database of case studies from a number of classes of slope (and dam) and material type.
The database included some 450 landslides in cuts, fills and natural slopes, and some
170 embankment dams.
For landslides in soil slopes, methods and guidelines have been developed for use in
the analysis, evaluation and prediction of the pre and post failure deformation
behaviour. They take into consideration the factors influencing and the mechanics
controlling the deformation behaviour for the classes of slope and material types, which
are different for pre and post failure. Pre-failure deformations are largely controlled by
the effective stress conditions within the slope, changes in the boundary conditions and
the response of the soil to those changes in boundary conditions. Whether the soil,
under the effective stress conditions imposed within the slope, is contractive (and
saturated or near saturated) or dilative on shearing, has a significant influence on the pre
failure deformation behaviour. The post failure deformation behaviour is strongly
influenced by the mechanics of failure (including whether the soil is contractive or
dilative on shearing), the source area slope angle, the downslope geometry, the
orientation of the surface of rupture, the material properties and slide volume.
Guidelines are presented for prediction of “rapid” and “slow” post failure velocity.
For embankment dams, methods and guidelines have been developed for evaluation
and prediction of the deformation behaviour during and post construction for selected
embankment types. They take into consideration the influence of material type and
placement methods, material strength and compressibility properties, embankment
zoning geometry, embankment height, and reservoir operation, amongst other factors.
Guidelines have been developed to assist in the identification of “abnormal”
deformation behaviour, which can be related to internal deformations or a marginal
stability condition and the onset to failure.
Acknowledgements Page ii

Acknowledgements
This project would not have been possible were it not for the vision and enthusiasm of
Professor Robin Fell. It is with heartfelt thanks and much appreciation that I
acknowledge Robin for his support, guidance and encouragement throughout the
project. It has been a most rewarding experience.
The support of the Australian Research Council and the industry sponsors of the
project (listed below) are acknowledged. I would like to thank those organizations, in
particular the personnel within those organizations, that gave of their time and efforts in
providing case study information, and feedback and discussion on aspects of the work.
I would also like to thank the staff of and the services provided by the School of
Civil and Environmental Engineering of the University of New South Wales. In
particular, I thank Messrs Paul Gwynne and Lindsay O’Keeffe for their efforts in the
establishment and the on-going running of the laboratory creep testing facility.
The friendship and support of my colleagues within the School of Civil and
Environmental Engineering is acknowledged. In particular I thank Doctor James
Glastonbury for his friendship, his listening and his thoughtful discussion. James and I
travelled on similar paths within the research project, and it has been a pleasure to have
made this journey with James.
Lastly, but by no means least, I would like to thank my friends and family for their
patience, encouragement and unquestioning support throughout the project.

Financial sponsors of the project:


• Australian Capital Territory Energy and Water Corporation;
• BC Hydro and Power Authority, Canada;
• Dams Safety Committee of New South Wales;
• DamWatch Services (New Zealand);
• Goulburn Murray Water;
• Gutteridge Haskins and Davey;
• Hydro Tasmania;
• Melbourne Water Corporation;
• Natural Resources and Environment, Victoria;
• New South Wales Department of Land and Water Conservation;
• New South Wales Department of Public Works and Services;
Acknowledgements Page iii

• Pacific Power;
• Pells Sullivan Meynink;
• Queensland Department of Main Roads;
• Roads and Traffic Authority of New South Wales;
• Snowy Mountains Engineering Corporation;
• Snowy Mountains Hydro-Electric Authority;
• South Australian Water Corporation;
• Sun Water (formerly Queensland Department of Natural Resources);
• Sydney Catchment Authority;
• United States Bureau of Reclamation;
• URS Corporation;
• Vattenfall, Sweden;
• Water Corporation of Western Australia.

In kind sponsors of the project:


• Australian Soil Testing;
• Geo-Eng Group;
• Geotechnical Engineering Office, Hong Kong Government;
• Jeffrey and Katauskas;
• River Murray Water;
• Wimmera Mallee Water.
Table of Contents Page iv

TABLE OF CONTENTS

1.0 INTRODUCTION ....................................................................... 1.1


1.1 OBJECTIVES OF THE R ESEARCH ....................................................................... 1.1
1.1.1 Landslides in Soil Slopes ........................................................................... 1.2
1.1.2 Deformation Behaviour of Embankment Dams......................................... 1.3
1.2 STRUCTURE OF THE THESIS ............................................................................. 1.4
1.3 TERMS AND DEFINITIONS ................................................................................ 1.5
1.3.1 General Terms and Definitions ................................................................. 1.5
1.3.2 Terms and Definitions for Landslides ....................................................... 1.5
1.3.3 Terms and Definitions for Embankments .................................................. 1.9

2.0 LITERATURE REVIEW ........................................................... 2.1


2.1 LANDSLIDES IN SOIL S LOPES ........................................................................... 2.1
2.2 DEFORMATION BEHAVIOUR OF EMBANKMENT DAMS ..................................... 2.7
2.3 WHERE T HIS R ESEARCH F ITS IN ......................................................................2.9

3.0 “RAPID” LANDSLIDES FROM FAILURES IN SOIL


SLOPES ....................................................................................... 3.1
3.1 OUTLINE OF THIS C HAPTER ............................................................................. 3.1
3.2 DATABASE OF CASE S TUDIES ANALYSED ........................................................ 3.2
3.2.1 Flow Slides in Coal Mine Waste Spoil Piles and Stockpiled Coal............ 3.5
3.2.2 Landslides from Failures in Cut, Fill and Natural Slopes, Hong
Kong .......................................................................................................... 3.7
3.2.3 Flow Slides from Tailings Dam Failures .................................................. 3.9
3.2.4 Flow Slides from Failures in Hydraulic Fill Embankment Dams............. 3.9
3.2.5 Flow Slides from Failures in Sub-Aqueous Constructed Fills.................. 3.9
3.2.6 Flow Slides from Failures in Submarine Slopes ..................................... 3.12
3.2.7 Flow Slides in Sensitive Clays................................................................. 3.12
3.3 THE MECHANICS OF SHEARING OF CONTRACTIVE GRANULAR SOILS ............ 3.15
3.3.1 Definitions ...............................................................................................3.15
3.3.2 Characteristics of Contractive and Dilative Soils Sheared Under
Monotonic Load Conditions.................................................................... 3.16
3.3.3 Field Methods for Evaluation of Flow Liquefaction Potential ...............3.22
Table of Contents Page v

3.3.4 Effects of Non-plastic Fines and Gravels based on Flow Liquefaction


Potential .................................................................................................. 3.30
3.3.5 Residual Undrained Shear Strength of Flow Liquefied Soils.................. 3.32
3.4 MECHANICS OF FAILURE OF FLOW SLIDES .................................................... 3.39
3.4.1 Mechanics of Development of Flow Sliding............................................ 3.39
3.4.2 Flow Slides in Granular Stockpiles......................................................... 3.43
3.4.3 Flow Slides from Failures in Sensitive Clays.......................................... 3.49
3.4.4 Flow Slides in Loose Silty Sand Fills, Hong Kong.................................. 3.54
3.4.5 Tailings Dams..........................................................................................3.57
3.4.6 Hydraulic Fill Embankment Dams..........................................................3.60
3.4.7 Submarine Slopes and Slopes in Sub-Aqueous Fills ............................... 3.62
3.5 MECHANICS OF FAILURE IN DILATIVE SOILS LEADING TO “RAPID”
SLIDING .........................................................................................................3.64
3.5.1 Slope Types of Failures in Dilative Soils Leading to “rapid” Sliding.... 3.64
3.5.2 Slides Through Soil Mass – Dilative Failures......................................... 3.65
3.5.3 Defect Controlled Landslides.................................................................. 3.79
3.6 SUMMARY OF THE CHARACTERISTICS AND MATERIAL T YPES OF “RAPID”
LANDSLIDES .................................................................................................. 3.83
3.6.1 Flow Slides in Contractant Soils Susceptible to Flow Liquefaction ....... 3.83
3.6.2 Characteristics and Identification of Conditions Under Which
Failures in Dilative Soils Result in “Rapid” Landsliding ...................... 3.88
3.7 PRE-FAILURE DEFORMATION BEHAVIOUR ...................................................... 3.89
3.7.1 Flow Slides .............................................................................................. 3.90
3.7.2 “Rapid” Landslides from Slope Failures in Dilative Soils ..................... 3.97
3.7.3 Summary of Pre Failure Deformation Behaviour ................................... 3.99
3.8 POST-FAILURE DEFORMATION BEHAVIOUR – R EVIEW OF THE
LITERATURE ................................................................................................3.102
3.8.1 Factors Affecting Travel Distance and Velocity ...................................3.102
3.8.2 Velocity of the Slide Mass .....................................................................3.104
3.8.3 Empirical Methods for Assessment of Travel Distance.........................3.107
3.9 POST-FAILURE DEFORMATION – DEVELOPED METHODS FOR ASSESSMENT
OF TRAVEL DISTANCE FROM THE D ATABASE ..............................................3.112
3.9.1 Travel Distance Versus Slide Volume and Initial Slide Classification .3.113
3.9.2 Travel Distance Versus Slide Volume, Degree of Confinement of the
Travel Path and Initial Slide Classification..........................................3.116
3.9.3 Travel Distance Versus Volume, Down-slope Angle and Slide Type
from Failures in Dilative Soils ..............................................................3.124
Table of Contents Page vi

3.9.4 Travel Distance Versus Volume, Down-slope Angle and Slide Type
from Failures in Contractile Soils.........................................................3.137
3.9.5 Summary of Methods for Predicting Travel Distances .........................3.150
3.10 POST-FAILURE NUMERICAL MODELLING - METHODS AND RESULTS ...........3.152
3.10.1 “Rapid” Landsliding in Hong Kong .....................................................3.155
3.10.2 “Rapid” Landslides from Coal Mine Waste Spoil Pile Failures in
British Columbia ...................................................................................3.158
3.10.3 Flow Slides from Tailings Dams Failures.............................................3.160
3.11 CONCLUSIONS .............................................................................................3.162

4.0 EMBANKMENTS ON SOFT GROUND................................... 4.1


4.1 INTRODUCTION TO THIS C HAPTER ................................................................... 4.1
4.2 LITERATURE R EVIEW ...................................................................................... 4.3
4.2.1 Excess Pore Water Pressure Response..................................................... 4.5
4.2.2 Deformation Behaviour...........................................................................4.15
4.3 ANALYSIS OF THE B EHAVIOUR OF FAILURE CASE STUDIES ........................... 4.22
4.3.1 Case Studies Analysed............................................................................. 4.22
4.3.2 Excess Pore Water Pressure ................................................................... 4.26
4.3.3 Pre-Failure Deformation Behaviour and the Effects of Progressive
Failure..................................................................................................... 4.32
4.3.4 Factor of Safety Versus Relative Embankment Height............................ 4.45
4.3.5 Post-Failure Deformation Behaviour ..................................................... 4.47
4.4 POST CONSTRUCTION BEHAVIOUR OF EMBANKMENTS ON SOFT GROUND .... 4.51
4.4.1 Effect of Effective Stress State on the Post Construction Behaviour....... 4.52
4.4.2 Deformation Behaviour and Pore Water Pressure Response in the
Initial Period Post Construction ............................................................. 4.52
4.4.3 Long-Term Post Construction Deformation............................................ 4.59
4.5 DISCUSSION AND CONCLUSIONS.................................................................... 4.61
4.5.1 Indicators of an Impending Failure ........................................................ 4.61
4.5.2 Guidelines on Monitoring for Identification of an Impending Failure
Condition................................................................................................. 4.63
4.5.3 Post Failure Deformation Behaviour...................................................... 4.64
4.5.4 Discussion on Failure Mechanism.......................................................... 4.65
Table of Contents Page vii

5.0 LANDSLIDES IN EMBANKMENT DAMS AND CUT


SLOPES IN HEAVILY OVER-CONSOLIDATED HIGH
PLASTICITY CLAYS ................................................................ 5.1
5.1 INTRODUCTION TO THIS C HAPTER ................................................................... 5.1
5.2 LITERATURE R EVIEW ...................................................................................... 5.2
5.2.1 Statistics on Dam Incidents Involving Slope Instability............................ 5.2
5.2.2 Progressive Failure...................................................................................5.8
5.2.3 Mechanics of Post Failure Deformation................................................. 5.11
5.3 DATABASE OF CASE S TUDIES ANALYSED ......................................................5.12
5.3.1 Case Studies of Failures in Dam Embankments ..................................... 5.14
5.3.2 Case Studies of Failures in Cut Slopes of High Plasticity Clays ............ 5.15
5.4 MECHANICS OF FAILURE OF C UT S LOPES IN HEAVILY OVER-
CONSOLIDATED HIGH PLASTICITY C LAYS ..................................................... 5.18
5.4.1 Progressive Failure................................................................................. 5.18
5.4.2 Trigger to Failure....................................................................................5.20
5.4.3 Time to Failure........................................................................................5.22
5.4.4 Effect of Defects Within the Soil Mass .................................................... 5.23
5.5 MECHANICS OF FAILURE FOR FAILURES IN EMBANKMENT DAMS .................5.24
5.5.1 Failures in Embankment Dams During Construction............................. 5.24
5.5.2 Failures in Embankments During Drawdown.........................................5.30
5.5.3 Post Construction Failures in the Downstream Slope of
Embankments...........................................................................................5.39
5.6 POST FAILURE DEFORMATION BEHAVIOUR – F AILURES IN EMBANKMENT
DAMS ............................................................................................................ 5.44
5.6.1 Factors Affecting the Post Failure Deformation..................................... 5.44
5.6.2 Summary of Case Studies of Failures in Embankment Dams ................. 5.45
5.6.3 Failures During Construction - Post Failure Deformation
Behaviour ................................................................................................ 5.46
5.6.4 Failures During Drawdown – Post Failure Deformation Behaviour.....5.68
5.6.5 Post Failure Deformation Behaviour of Failures in the Downstream
Shoulder After Construction....................................................................5.80
5.7 POST FAILURE DEFORMATION B EHAVIOUR - C UT S LOPES IN HEAVILY
OVER-CONSOLIDATED C LAYS ....................................................................... 5.86
5.7.1 Summary of Factors Affecting the Post Failure Deformation ................ 5.86
5.7.2 Summary of the Post Failure Deformation and Velocity of the
Failure Case Studies ............................................................................... 5.87
Table of Contents Page viii

5.7.3 Failures in Cut Slopes – Type 2 Slope Failure Geometry....................... 5.92


5.7.4 Failures in Cut Slopes – Type 1 Slope Failure Geometry....................... 5.94
5.7.5 Failures in Cut Slopes – Type 5 Slope Failure Geometry....................... 5.95
5.8 PREDICTION OF POST-FAILURE TRAVEL DISTANCE ....................................... 5.97
5.8.1 Khalili et al (1996) Model for Intact Slides ............................................ 5.97
5.8.2 Empirical Methods for Prediction of Travel Distance of Intact Slides.5.105
5.9 CONCLUSIONS AND GUIDELINES FOR PREDICTION OF POST-FAILURE
DEFORMATION BEHAVIOUR OF INTACT SLIDES IN SOIL SLOPES ..................5.108
5.9.1 Summary of Findings from the Case Study Analysis.............................5.108
5.9.2 Guidelines for Prediction of Post-Failure Deformation.......................5.117

6.0 THE DEFORMATION BEHAVIOUR OF ROCKFILL .......... 6.1


6.1 OUTLINE OF THIS CHAPTER .............................................................................6.1
6.2 DEFINITIONS AND TERMINOLOGY ....................................................................6.2
6.2.1 Unconfined Compressive Strength of Rock............................................... 6.2
6.2.2 Rockfill Placement and Compaction ......................................................... 6.3
6.2.3 Rockfill Moduli..........................................................................................6.4
6.2.4 Zoning of Main Rockfill in Concrete Face Rockfill Dams ........................ 6.7
6.3 LITERATURE R EVIEW OF ROCKFILL DEFORMATION ........................................ 6.9
6.3.1 Historical Summary of Rockfill Usage in Embankment Design................ 6.9
6.3.2 Deformation Properties of Rockfill ........................................................... 6.9
6.3.3 Predictive Methods for Rockfill Deformation ......................................... 6.14
6.4 ANALYSIS OF THE DEFORMATION BEHAVIOUR OF ROCKFILL IN CONCRETE
FACE ROCKFILL DAMS .................................................................................. 6.27
6.4.1 Case Study Database...............................................................................6.28
6.4.2 Deformation During Construction .......................................................... 6.28
6.4.3 Deformation of the Face Slab of CFRD on First Filling ........................ 6.44
6.4.4 Post Construction Crest Settlement......................................................... 6.53
6.5 DISCUSSION AND RECOMMENDED METHODS FOR PREDICTION ..................... 6.71
6.5.1 Guidelines on Deformation Prediction During Construction................. 6.71
6.5.2 Guidelines on Deformation Prediction During First Filling .................. 6.75
6.5.3 Guidelines on Deformation Prediction Post-Construction..................... 6.76
6.6 CONCLUSIONS ............................................................................................... 6.77
Table of Contents Page ix

VOLUME 2

7.0 THE DEFORMATION BEHAVIOUR OF


EMBANKMENT DAMS............................................................. 7.1
7.1 INTRODUCTION TO THIS C HAPTER ................................................................... 7.1
7.2 LITERATURE R EVIEW ...................................................................................... 7.2
7.2.1 Failure and Accident Statistics.................................................................. 7.2
7.2.2 Historical Development of Embankment Dams as this Affects
Deformation Behaviour.............................................................................7.3
7.2.3 Factors Affecting The Deformation of Embankment Dams ...................... 7.5
7.2.4 Predictive Methods of Deformation Behaviour......................................... 7.9
7.3 DATABASE OF CASE S TUDIES ........................................................................7.17
7.3.1 Earthfill and Zoned Earth and Earth-Rockfill Embankments................. 7.17
7.3.2 Puddle Core Earthfill Embankments....................................................... 7.18
7.4 GENERAL DEFORMATION B EHAVIOUR DURING CONSTRUCTION OF

EARTHFILL AND ZONED E ARTH AND EARTH-ROCKFILL EMBANKMENTS ...... 7.18


7.4.1 Stresses During Construction.................................................................. 7.23
7.4.2 Lateral Deformation of the Core During Construction .......................... 7.43
7.4.3 Vertical Deformation of the Core During Construction......................... 7.51
7.5 GENERAL DEFORMATION BEHAVIOUR ON FIRST F ILLING OF EARTHFILL
AND ZONED EARTH AND E ARTH-ROCKFILL EMBANKMENTS ......................... 7.69
7.5.1 Effect of Water Load on the Core............................................................ 7.70
7.5.2 Collapse Compression During First Filling ........................................... 7.73
7.5.3 Lateral Surface Deformation Normal to the Dam Axis During First
Filling. ..................................................................................................... 7.84
7.6 GENERAL POST CONSTRUCTION DEFORMATION BEHAVIOUR OF

EARTHFILL AND ZONED E ARTH AND EARTH-ROCKFILL EMBANKMENTS ...... 7.99


7.6.1 Post Construction Internal Vertical Deformation of the Core..............7.100
7.6.2 Post Construction Total Deformation of Surface Monitoring Points ...7.102
7.6.3 Post Construction Crest Settlement Versus Time..................................7.110
7.6.4 Post Construction Horizontal Displacement of the Crest Normal to
the Dam Axis. ........................................................................................7.132
7.6.5 Post Construction Deformation of the Mid to Upper Downstream
Slope ......................................................................................................7.140
Table of Contents Page x

7.6.6 Post Construction Deformation of the Upper Upstream Slope and


Upstream Crest......................................................................................7.148
7.7 GENERAL DEFORMATION BEHAVIOUR OF PUDDLE CORE EARTHFILL
EMBANKMENTS ...........................................................................................7.158
7.7.1 Deformation During Construction of Puddle Core Earthfill
Embankments.........................................................................................7.158
7.7.2 Deformation During First Filling of Puddle Core Earthfill
Embankments.........................................................................................7.159
7.7.3 Deformation Behaviour Post First Filling of Puddle Core Earthfill
Dams......................................................................................................7.166
7.8 “ABNORMAL” EMBANKMENT DEFORMATION BEHAVIOUR – METHODS OF
IDENTIFICATION ..........................................................................................7.185
7.9 “ABNORMAL” DEFORMATION BEHAVIOUR DURING CONSTRUCTION OF
EARTH AND EARTH-ROCKFILL EMBANKMENTS ..........................................7.187
7.9.1 Plastic Deformation of the Core During Construction.........................7.188
7.9.2 Collapse Compression of the Central Earthfill Zone............................7.191
7.9.3 Reservoir Filling During Construction .................................................7.192
7.9.4 Shear Surface Development in the Core During Construction.............7.193
7.10 “ABNORMAL” DEFORMATION BEHAVIOUR POST CONSTRUCTION OF
ZONED E ARTH AND ROCKFILL EMBANKMENTS ...........................................7.196
7.10.1 Rockfill Susceptibility to Collapse Compression...................................7.197
7.10.2 Deformation on First Filling in Embankments where Collapse
Compression occurs in the Upstream Rockfill Shoulder.......................7.201
7.10.3 Development of Shear Surfaces Within the Earthfill Core....................7.217
7.10.4 Post First Filling Acceleration in Deformation that is Not Known to
be Shear Related....................................................................................7.241
7.10.5 Other Case Studies with Potentially “Abnormal” Deformation
Behaviour Post Construction ................................................................7.243
7.11 “ABNORMAL” DEFORMATION BEHAVIOUR POST CONSTRUCTION OF
EARTHFILL AND ZONED EMBANKMENTS WITH VERY BROAD CORE
WIDTHS .......................................................................................................7.252
7.11.1 Collapse Compression of the Earthfill on Wetting................................7.252
7.11.2 “High” Shear Stress or Marginal Stability Conditions Within the
Embankment ..........................................................................................7.260
7.11.3 The Effect of the Development of the Phreatic Surface on the
Displacement of the Crest and Downstream Shoulder. ........................7.269
7.11.4 Other Cases of Potential “Abnormal” Deformation Behaviour...........7.270
Table of Contents Page xi

7.12 “ABNORMAL” DEFORMATION BEHAVIOUR OF PUDDLE CORE EARTHFILL


DAMS ..........................................................................................................7.274
7.12.1 Comparison with the Deformation Behaviour of Other Puddle Dams .7.275
7.12.2 General Movement Trends Indicative of Deformation to Failure.........7.278
7.12.3 Other Indicators of “Abnormal” Deformation Behaviour....................7.279
7.13 SUMMARY OF “ABNORMAL” DEFORMATION BEHAVIOUR ...........................7.280
7.14 SUMMARY AND METHODS FOR PREDICTION OF DEFORMATION OF

EMBANKMENT DAMS ..................................................................................7.287


7.14.1 Earthfill, and Zoned Earth and Earth-Rockfill Dams ...........................7.288
7.14.2 Prediction of Deformation Behaviour During Construction for
Earthfill and Zoned Earth and Earth-Rockfill Embankments...............7.288
7.14.3 Prediction of Deformation Behaviour Post Construction for Earthfill
and Zoned Earth and Earth-Rockfill Embankments..............................7.291
7.14.4 Puddle Core Earthfill Dams..................................................................7.300
7.15 CONCLUSIONS .............................................................................................7.302

8.0 CONCLUSIONS AND RECOMMENDATIONS...................... 8.1


8.1 CONCLUSIONS ................................................................................................. 8.1
8.1.1 Landslides in Soil Slopes ...........................................................................8.1
8.1.2 Deformation Behaviour of Embankment Dams......................................... 8.4
8.2 RECOMMENDATIONS FOR FURTHER R ESEARCH ............................................... 8.6

9.0 REFERENCES ............................................................................ R1

VOLUME 3

APPENDICES
List of Tables Page xii

LIST OF TABLES
Table 1.1: IUGS (1995) velocity classifications for landslides. .................................... 1.6
Table 1.2: Classification of unconfined compressive strength of intact rock (AS
1726-1993) ............................................................................................................ 1.13

Table 3.1: Summary of database on “rapid” landslides. ................................................ 3.3


Table 3.2: Summary of data sources for “rapid” landslides from coal mine waste
spoil piles and coal stockpiles................................................................................. 3.5
Table 3.3: Summary of landslide cases analysed from Hong Kong .............................. 3.8
Table 3.4: Summary of case studies of “rapid” landslides from tailings dams ............ 3.10
Table 3.5: Summary of failure case studies in hydraulic fill embankment dams ........ 3.11
Table 3.6: Summary of failure case studies in sub-aqueous constructed fill slopes .... 3.11
Table 3.7: Summary of “rapid” flow slides in natural submarine slopes..................... 3.13
Table 3.8: Summary of flow slides in sensitive clays .................................................. 3.14
Table 3.9: SPT blow count ((N1)60) yield strength and residual strength correction
factors for fines content (Seed et al 1985; Seed 1987)..........................................3.35
Table 3.10: Features of flow slides in loose fill from Hong Kong...............................3.56
Table 3.11: Summary of properties of liquefaction susceptible materials from flow
slides in hydraulic fill embankments..................................................................... 3.61
Table 3.12: Summary of material types of dilative slides of debris in natural
slopes that developed in “rapid” debris slides and debris flows. ..........................3.75
Table 3.13: Source area slope angle ( α 1 ) of dilative slides in natural slopes that
developed into “rapid” debris flows and debris slides..........................................3.77
Table 3.14: Failures through the soil mass in dilative soils that developed into
“rapid” landslides, from Hong Kong..................................................................... 3.78
Table 3.15: Features of defect controlled compound slides that developed into
“rapid” landslides from Hong Kong...................................................................... 3.82
Table 3.16: Features of defect controlled translational slides that developed into
“rapid” landslides from Hong Kong...................................................................... 3.82
Table 3.17: Summary of pre-failure observations in coarse-grained loose fills .......... 3.92
Table 3.18: Measured slope deformations adjacent to failures in the Ottawa area
(Mitchell and Eden 1972)......................................................................................3.95
Table 3.19: Summary of characteristics of pre-failure deformation of slope
failures that developed into “rapid” landslides ...................................................3.100
List of Tables Page xiii

Table 3.20: Velocity of the slide mass for “rapid” landslides from failures in soil
slopes (mostly from case studies analysed).........................................................3.105
Table 3.21: Velocity of slide mass for debris slides and debris flows.......................3.106
Table 3.22: Regression analysis of H/L (Equation 3.11) versus landslide volume
(Corominas 1996a)..............................................................................................3.111
Table 3.23: Summary of empirical correlation coefficients (Equation 3.11) for
debris flows by Corominas (1996a) ....................................................................3.114
Table 3.24: Empirical correlation coefficients for power law fit of travel distance
angle to slide volume (Equation 3.11, H L = AV B ) for unconfined “rapid”
landslides.............................................................................................................3.122
Table 3.25: Empirical correlation coefficients for power law fit of travel distance
angle to slide volume (Equation 3.11, H L = AV B ) for confined and partly
confined “rapid” landslides.................................................................................3.124
Table 3.26: Summary of empirical correlations for flow slides in coal waste spoil
piles in British Columbia ....................................................................................3.139
Table 3.27: Statistical summary of H/L to tan a 2 correlation for flow slides in
coarse-grained coal mine waste spoil piles.........................................................3.143
Table 3.28: Summary of slide properties of flow slides in hydraulic fill
embankments.......................................................................................................3.149
Table 3.29: Mean and standard deviation of H/L for several types of slopes giving
“rapid” landslides................................................................................................3.153
Table 3.30: Summary of recommended methods for prediction of H/L (tangent of
the travel distance angle).....................................................................................3.154
Table 3.31: Summary of “rapid” landslides from Hong Kong analysed by Hungr
Geotechnical Research (1998) and Ayotte and Hungr (1998) ............................3.156
Table 3.32: Summary of results of numerical modelling using DAN for “rapid”
landslides in Hong Kong (Hungr Geotechnical Research 1998; Ayotte and
Hungr 1998) ........................................................................................................3.157

Table 4.1: Summary of failure case studies of embankments on soft ground. ............ 4.24
Table 4.2: Excess pore water pressure observations for the failure case studies......... 4.26
Table 4.3: Summary of exceptions to general observations of excess pore water
pressure response. ................................................................................................. 4.31
Table 4.4: Summary of case studies that are exceptions or have other explanations
to deformation as an indicator of impending failure............................................. 4.44
Table 4.5: Results of the post-failure deformation analysis......................................... 4.50
List of Tables Page xiv

Table 5.1: Incidence of slope instability in embankment dams (Foster 1999) .............. 5.3
Table 5.2: Factors influencing downstream slides - accident and failure incidents
(Foster 1999) ........................................................................................................... 5.4
Table 5.3: Factors influencing upstream slides - accident and failure incidents
(Foster 1999) ........................................................................................................... 5.5
Table 5.4: Failures involving slope instability, excluding sloughing cases (Foster
1999) ....................................................................................................................... 5.9
Table 5.5: Summary of case study database of slope failures...................................... 5.13
Table 5.6: Summary of embankment dam slope stability incidents by failure slope
geometry................................................................................................................ 5.15
Table 5.7: Slope instability case studies in embankment dams by embankment
type........................................................................................................................ 5.16
Table 5.8: Mechanics of failure for failures in embankment dams during
construction. .......................................................................................................... 5.24
Table 5.9: Material types through which the surface of rupture passed of post
construction failures in the downstream shoulder................................................. 5.40
Table 5.10: Timing of the failure for post construction slides in the downstream
shoulder. ................................................................................................................ 5.42
Table 5.11: Distribution of slide volume for slides in embankment dams. ................. 5.46
Table 5.12: Summary of failure case studies in embankment dams during
construction – failures within the embankment only ............................................ 5.47
Table 5.13: Summary of failure case studies in embankment dams during
construction – failure within the embankment and foundation............................. 5.48
Table 5.14: Summary of the case studies of slides in the upstream shoulder of
embankment dams triggered by drawdown .......................................................... 5.69
Table 5.15: Post failure deformation behaviour and slope/material properties of
slides in embankment dams during drawdown. .................................................... 5.72
Table 5.16: Summary of the case studies of slides in the downstream shoulder of
embankment dams that occurred post construction. ............................................. 5.82
Table 5.17: Cut slope failures in London clay. ............................................................ 5.88
Table 5.18: Cut slope failures in Upper Lias clay........................................................ 5.89
Table 5.19: Results of the post-failure deformation analysis using the Khalili et al
(1996) method. ....................................................................................................5.102
Table 5.20: Characteristics of the case study groupings from the post failure
deformation analysis (Figure 5.52) .....................................................................5.104
Table 5.21: Factors affecting the peak post-failure velocity for slides in
embankment dams during drawdown. ................................................................5.116
List of Tables Page xv

Table 5.22: Guidelines for post-failure deformation prediction using the Khalili et
al (1996) models..................................................................................................5.118

Table 6.1: Classification of unconfined compressive strength or rock (AS 1726-


1993) ....................................................................................................................... 6.2
Table 6.2: Historical summary of rockfill usage in embankment design (Galloway
1939; Cooke 1984; Cooke 1993). ......................................................................... 6.10
Table 6.3: Parameters for deformation prediction (Soydemir and Kjærnsli 1979)...... 6.25
Table 6.4: Rates of post-construction crest settlement of dumped and compacted
rockfills in CFRDs (Sherard and Cooke 1987) ..................................................... 6.26
Table 6.5: Summary of embankment and rockfill properties for CFRD case
studies.................................................................................................................... 6.29
Table 6.6: Assessment of cross-valley influence on arching for case studies
analysed................................................................................................................. 6.38
Table 6.7: Stress conditions representative of the data sets from Figure 6.21............. 6.43
Table 6.8: Summary of crest settlement during the period of first filling.................... 6.62
Table 6.9: Values of the coefficient, m, in the strain rate – time power function
(Equation 6.5)........................................................................................................ 6.63
Table 6.10: Estimates of long-term crest settlement rates for dumped rockfills.......... 6.67
Table 6.11: Location of internal post-construction vertical settlement in CFRD ........ 6.70
Table 6.12: Approximate stress reduction factors to account for valley shape............ 6.73

Table 7.1: Mechanisms affecting the long-term deformation behaviour of old


puddle core earthfill embankments (Tedd et al 1997a)........................................... 7.8
Table 7.2: Vertical compression of rolled, well-compacted earthfills measured
during construction (adapted from Sherard et al 1963)......................................... 7.12
Table 7.3: Published ranges of post construction deformation of embankment
dams ...................................................................................................................... 7.14
Table 7.4: Central core earth and rockfill embankments in the database .................... 7.19
Table 7.5: Zoned earth and rockfill embankment case studies .................................... 7.20
Table 7.6: Zoned earthfill embankment case studies ................................................... 7.21
Table 7.7: Earthfill embankment case studies.............................................................. 7.21
Table 7.8: Puddle core earthfill embankment case studies .......................................... 7.22
Table 7.9: Multiplying coefficients for the calculation of the total vertical stress at
the embankment centreline at end of construction (Equation 7.3)........................ 7.26
Table 7.10: Properties of central core used in finite difference analyses..................... 7.29
List of Tables Page xvi

Table 7.11: Lateral deformations of the central core at end of construction for
central core earth and rockfill embankments. ....................................................... 7.44
Table 7.12: Summary of embankment and earthfill properties for cases used in the
analysis of lateral core deformation during construction. ..................................... 7.46
Table 7.13: Confined secant moduli during construction for well-compacted, dry
placed earthfills. .................................................................................................... 7.57
Table 7.14: Equations of best fit for core settlement versus embankment height
during construction................................................................................................ 7.68
Table 7.15: Equations of best fit for core settlement (as a percentage of
embankment height) versus embankment height during construction.................. 7.68
Table 7.16: Properties of central core used in finite difference analyses..................... 7.80
Table 7.17: Lateral displacement of the crest (centre to downstream edge) on first
filling. .................................................................................................................... 7.89
Table 7.18: Thirteen cases with largest downstream crest displacement on first
filling ..................................................................................................................... 7.92
Table 7.19: Typical range of post construction crest settlement. ...............................7.104
Table 7.20: Typical range of post construction settleme nt of the upstream and
downstream shoulders.........................................................................................7.109
Table 7.21: Summary of long-term crest settlement rates (% per log cycle of
time). ...................................................................................................................7.122
Table 7.22: Range of long-term settlement rate of the downstream shoulder. ..........7.141
Table 7.23: Range of long-term settlement rate of the upper upstream slope and
upstream crest region. .........................................................................................7.151
Table 7.24: Summary of the post construction surface deformations of the puddle
core earthfill dam case studies. ...........................................................................7.168
Table 7.25: Embankments for which collapse compression caused moderate to
large settlements..................................................................................................7.199
Table 7.26: Figure references for post construction crest deformation......................7.292
Table 7.27: Figure references for post construction deformation of the
embankment shoulders........................................................................................7.293
Table 7.28: References to post construction settlement magnitude tables and plots .7.295
Table 7.29: Embankment crest region, typical range of post construction
settlement and long-term settlement rate ............................................................7.295
Table 7.30: Embankment shoulder regions, typical range of post construction
settlement and long-term settlement rate ............................................................7.296
Table 7.31: References to tables and figures of long-term settlement rate ................7.296
List of Tables Page xvii

Table 7.32: Predictive methods of long-term crest settlement and displacement


under normal reservoir operating conditions for puddle core earthfill
embankments.......................................................................................................7.301
List of Figures Page xviii

LIST OF FIGURES
Figure 1.1: Landslide classification system and main slide types.................................. 1.7
Figure 1.2: Definition of travel distance, travel distance angle, and slope
geometry. ................................................................................................................. 1.8
Figure 1.3: Slope failure geometries (a) Type 1 - failure at the top of the cut or fill
slope, (b) Type 2 – the toe of failure is coincident with the toe of the slope,
and (c) Type 5 – the surface of rupture extends below and daylights beyond
the toe of the slope. ................................................................................................. 1.8
Figure 1.4: Classification of slope failure geometries (a) Type 3 –travel on
relatively uniform or gradually decreasing slope angles, and (b) Type 4 –
travel on changing slope angles from steep to shallow...........................................1.9
Figure 1.5: Embankment zoning classification system (Foster 1999) .........................1.11
Figure 1.6: Dam zoning categories of embankment types (Foster et al 2000)............. 1.11

Figure 2.1: Schematic of the system for geotechnical characterisation of slope


movement (Leroueil et al 1996)..............................................................................2.3
Figure 2.2: The stages of slope movement (Leroueil et al 1996)................................... 2.4
Figure 2.3: Concept of the creep model under constant deviatoric stress
conditions. ............................................................................................................... 2.6
Figure 2.4: Monitored deformation behaviour at retained cut slope failure in
London clay at Kensal Green (Skempton 1964) ..................................................... 2.7

Figure 3.1: Typical coal waste spoil pile profile in British Columbia (Hungr et al
1998) ....................................................................................................................... 3.6
Figure 3.2: Steady state line of Toyoura sand (Ishihara 1993) .................................... 3.16
Figure 3.3: Undrained behaviour of states on the contractive side of the steady
state line (Ishihara 1993)....................................................................................... 3.17
Figure 3.4: Undrained behaviour from states close to steady state and on dilative
side (a) stress-strain plot and (b) effective stress paths (Ishihara 1993) ............... 3.17
Figure 3.5: Schematic diagram showing different types of behaviour for undrained
stress paths of samples at the same void ratio and different initial confining
pressures (Yamamuro and Lade 1998).................................................................. 3.18
Figure 3.6: Definition of the state parameter (Been and Jefferies 1985). .................... 3.20
Figure 3.7: Collapse surface concept, defined by the locus of the peak deviator
stress from undrained triaxial samples tested at the same void ratio and
different initial effective mean normal stress (Sladen et al 1985a)....................... 3.20
List of Figures Page xix

Figure 3.8: Correlation for cyclic resistance ratio (CRR) based on corrected SPT
blow count, (N1)60, for clean sands under level ground conditions (Robertson
and Wride 1997).................................................................................................... 3.23
Figure 3.9: Corrected SPT blow count, (N1)60, versus vertical effective over-
′ ) for saturated non-gravelly silty sand deposits that have
burden pressure ( σ vo
experienced large deformations and were triggered by earthquake (Baziar and
Dobry 1995) .......................................................................................................... 3.24
Figure 3.10: Flow liquefaction and no-flow bounds in terms of (a) field measured
SPT blow count and (b) CPT qc value from flow liquefaction case histories
(static and earthquake triggered) and laboratory undrained testing (Ishihara
1993). .................................................................................................................... 3.25
Figure 3.11: Influence of void ratio range on flow potential of sandy soils for (a)
round-grained sands and (b) angular grains (Cubrinovski and Ishihara 2000)..... 3.26
Figure 3.12: Relationship between relative density, Dro , at the threshold void ratio,
eo, and void ratio range (emax – emin) for sandy soils (Cubrinovski and Ishihara
2000). .................................................................................................................... 3.27
Figure 3.13: Relationship between the slope of the steady state line, λ , and void
ratio range for (a) round-grained sands and (b) angular grained sands
(Cubrinovski and Ishihara 2000)........................................................................... 3.28
2
Figure 3.14: Correlation between N1 Dr (N1 = SPT blow count corrected for
over-burden pressure, Dr = relative density) and void ratio range for
cohesionless soils (Cubrinovski and Ishihara 2000). ............................................ 3.28
Figure 3.15: Field measured SPT blow count boundaries differentiating flow
liquefaction no-flow conditions for void ratio range of 0.3 to 0.6 for (a)
round-grained sands and (b) angular sands (Cubrinovski and Ishihara 2000)...... 3.29
Figure 3.16: Flow characterisation of sand based on field measured SPT blow
count showing liquefaction flow no-flow boundary, and region of flow at zero
residual undrained strength (Cubrinovski and Ishihara 2000) .............................. 3.30
Figure 3.17: Increase in liquefaction potential (as measured by relative density)
with increasing content of non-plastic fines (Lade and Yamamuro 1997) ........... 3.31
Figure 3.18: Fording sandy gravel sample (a) undrained triaxial compression test
results and (b) consolidation and steady states (Dawson et al 1998). ................... 3.32
Figure 3.19: Relationship between equivalent clean sand SPT blow count, (N1)60-
CS , and residual undrained strength from field case studies (Seed and Harder

1990) ..................................................................................................................... 3.34


Figure 3.20: Relationship between undrained critical strength ratio and equivalent
clean sand SPT blow count, (N1)60-CS (Stark and Mesri 1992).............................. 3.35
List of Figures Page xx

Figure 3.21: Residual undrained shear strength, Sr, to vertical effective over-
burden pressure for silty sand deposits (≥ 10% fines), triggered by
earthquake, that have experienced large deformation (Baziar and Dobry 1995).. 3.36
Figure 3.22: Flow chart for assessment of residual undrained shear strength of
liquefied soil (NSF 1998)...................................................................................... 3.38
Figure 3.23: Idealised mechanism for initiation of flow slides....................................3.40
Figure 3.24: Definition of undrained brittleness index, IB, in strain weakening soil. .. 3.41
Figure 3.25: Effective stress states on some planes (submarine slopes) located
within the region of potentially instability (Lade 1992). ......................................3.42
Figure 3.26: Collapse model for mine waste dumps (Dawson et al 1998) ..................3.43
Figure 3.27: Particle size distribution of coal mine waste and coking coal
susceptible to static liquefaction ........................................................................... 3.44
Figure 3.28: Particle size distribution of sensitive clays ..............................................3.50
Figure 3.29: Frequency distribution of landslides in Quebec (Tavenas 1984) ............3.52
Figure 3.30: Slope height versus slope inclination for unstable slopes in eastern
Canada (Tavenas 1984)......................................................................................... 3.52
Figure 3.31: Retrogressive landslide in sensitive clay (Leroueil et al 1996) ...............3.54
Figure 3.32: Peak shear strength from undrained triaxial compression tests of
contractive fill samples (HKIE 1998) ...................................................................3.57
Figure 3.33: Effect of partial saturation on steady state line in void ratio – mean
effective stress space (HKIE 1998)....................................................................... 3.58
Figure 3.34: Particle size distribution of tailings from flow slide case studies............3.59
Figure 3.35: Particle size distributions of hydraulic fills from flow slide case
studies.................................................................................................................... 3.61
Figure 3.36: Particle size distributions of soil types from submarine slopes and
sub-aqueous fills susceptible to liquefaction and flow sliding.............................. 3.62
Figure 3.37: Hypothetical model for transformation of dilative materials from
rigid to fluid behaviour, path Di to Df (Fleming et al 1989).................................. 3.68
Figure 3.38: Relationship between peak hourly rainfall, landsliding and severity of
consequence for Hong Kong (Brand et al 1984)................................................... 3.70
Figure 3.39: Relationship between severity of landsliding and rainfall intensity for
Hong Kong (Brand 1985b).................................................................................... 3.70
Figure 3.40: Source area particle size distributions of dilative slides of debris in
natural slopes......................................................................................................... 3.73
Figure 3.41: Distribution of head slope angle (source area slope angle) for slides
of debris in natural terrain that developed into “rapid” landslides, Hong Kong
(Evans et al 1997).................................................................................................. 3.79
List of Figures Page xxi

Figure 3.42: Cross section through the failure and “rapid” landslide at Fei Tsui
Road, Hong Kong (GEO 1996a)........................................................................... 3.81
Figure 3.43: Particle size distributions of material types susceptible to liquefaction
and flow sliding. .................................................................................................... 3.84
Figure 3.44: Comparison of flow liquefaction no-flow boundaries (in terms of
SPT (N1)60) for sands and silty sands from monotonic laboratory undrained
tests and earthquake triggered field cases (after Cubrinovski and Ishihara
2000). .................................................................................................................... 3.87
Figure 3.45: Pre-failure deformation of two separate failures at the Clode waste
dump, Fording Coal, British Columbia (Campbell and Shaw 1978).................... 3.93
Figure 3.46: Measured deformations in a natural slope, Ottawa River (Mitchell
and Eden 1972) ..................................................................................................... 3.94
Figure 3.47: Deformation and pore pressure ratio of an induced slope failure
(Mitchell and Williams 1981) ............................................................................... 3.96
Figure 3.48: Sliding block model showing kinetic energy lines for different
rheological models (Golder Assoc. 1995)...........................................................3.106
Figure 3.49: Schematic profile of a slope failure showing travel distance and
travel distance angle............................................................................................3.108
Figure 3.50: Travel distance from the toe of the cut, R, versus cut slope height, Hs,
for failures in cut slopes from Lantau Island, Hong Kong (Wong and Ho
1996) ...................................................................................................................3.109
Figure 3.51: Landslide volume versus tangent of the travel distance angle for 204
landslides (Corominas 1996a).............................................................................3.110
Figure 3.52: Ratio H/L versus volume for landslides that break-up on sliding
(Finlay et al 1999) ...............................................................................................3.111
Figure 3.53: Ratio H/L versus volume for all slides from database...........................3.114
Figure 3.54: H/L ratio versus volume relationships for Corominas (1996a) data on
unconfined debris flows. .....................................................................................3.115
Figure 3.55: H/L ratio versus volume relationships for Corominas (1996a) data on
confined debris flows. .........................................................................................3.115
Figure 3.56: Travel distance beyond source area versus volume for unconfined
travel paths. .........................................................................................................3.117
Figure 3.57: Travel distance beyond source area versus volume for confined and
partly confined travel paths. ................................................................................3.117
Figure 3.58: Comparison of trendlines of H/L versus slide volume differentiated
on travel path confinement for “rapid” landslides from failures in dilative
soils, from Hong Kong. .......................................................................................3.118
List of Figures Page xxii

Figure 3.59: H/L versus volume for “rapid” landslides with unconfined travel
paths. ...................................................................................................................3.120
Figure 3.60: H/L ratio versus volume for “rapid” landslides from failures in
dilative soils with unconfined travel paths..........................................................3.120
Figure 3.61: H/L ratio versus volume for “rapid” landslides from failures in
contractive soils (flow slides) with unconfined travel paths. ..............................3.121
Figure 3.62: H/L versus volume for “rapid” landslides with confined travel paths...3.123
Figure 3.63: H/L versus volume for “rapid” landslides with partly confined travel
paths. ...................................................................................................................3.123
Figure 3.64: H/L versus slide volume for cut slope cases studies from Hong Kong .3.126
Figure 3.65: H/L normalised against tangent of the cut slope angle versus volume
for cut slopes in Hong Kong failing onto a near horizontal slope at the toe.......3.127
Figure 3.66: Idealised effect of volume on travel distance for cut slope failures of
(a) small volume compared with (b) significantly larger volumes. ....................3.127
Figure 3.67: Classification of slope categories (a) Type 1 – failure at top of cut
slope, and (b) Type 2 – failure encompassing the full cut height. ......................3.128
Figure 3.68: Definition of down-slope angle below source area, a 2..........................3.132
Figure 3.69: H/L versus tangent of the down-slope angle, a 2, for unconfined travel
paths, failures in natural slopes from Hong Kong...............................................3.133
Figure 3.70: H/L versus tangent of the down-slope angle, a 2, for partly confined
travel paths, failures in natural slopes from Hong Kong.....................................3.134
Figure 3.71: H/L versus tangent of the down-slope angle, a 2, for confined travel
paths, failures in natural slopes from Hong Kong...............................................3.134
Figure 3.72: Natural slopes (Hong Kong). Trendlines for H/L versus tangent of
the down-slope angle, a 2, for all types of travel path. .........................................3.135
Figure 3.73: H/L versus slide volume for flow slides in loose silty sand to sandy
silt fills.................................................................................................................3.137
Figure 3.74: H/L versus slide volume for flow slides in coarse-grained coal mine
waste spoil piles. .................................................................................................3.142
Figure 3.75: H/L versus the tangent of the down-slope angle below the toe of the
spoil pile, a 2; flow slides in coarse-grained coal mine waste spoil piles. ...........3.142
Figure 3.76: H/L versus volume for retrogressive “rapid” slides...............................3.146
Figure 3.77: Distance of retrogression versus liquidity index for slides in sensitive
clays.....................................................................................................................3.148
Figure 3.78: Distance of retrogression based on stability number for landslides in
sensitive clays (Mitchell and Markell 1974).......................................................3.148
Figure 3.79: Ratio of retrogression distance to slide depth versus stability number .3.149
List of Figures Page xxiii

Figure 3.80: DAN results of frictional back-analysis models of “rapid” landslides


from Hong Kong (data from Hungr GR (1998) and Ayotte and Hungr (1998)) 3.158
Figure 3.81: Back calculated bulk friction angle from DAN analysis of flow slides
from coal mine waste spoil pile failures in British Columbia (Golder Assoc.
1995) ...................................................................................................................3.159
Figure 3.82: Simple sliding block model for analysis of tailings strength (Blight
1997) ...................................................................................................................3.160
Figure 3.83: Variation of shear strength with moisture content for gold tailings
(Blight 1997) .......................................................................................................3.161
Figure 3.84: Model for flow of liquefied tailings (Jeyapalan et al 1983a).................3.162

Figure 4.1: Computed factor of safety for field failures (Tavenas and Leroueil
1980) ....................................................................................................................... 4.2
Figure 4.2: Effective vertical stress profile and pre-consolidation stress profile
prior to construction, and the effective vertical stress profile at end of
construction of the Saint Alban B test embankment (Tavenas and Leroueil
1980). ...................................................................................................................... 4.8
Figure 4.3: Field observations of excess pore water pressures in clay foundations
during the initial period of embankment construction (Tavenas and Leroueil
1980). ...................................................................................................................... 4.8
Figure 4.4: Total (OPFR) and effective (O′P′F′R′) stress paths under the centre of
embankments (Tavenas and Leroueil 1980) ........................................................... 4.9
Figure 4.5: Relationship between the excess pore water pressure and the applied
vertical stress (Tavenas and Leroueil 1980)............................................................ 4.9
Figure 4.6: Dimensionless variations of changes in total stresses under the
centreline of an infinite strip load on an elastic half space (Poulos and Davis
1974) ..................................................................................................................... 4.10
Figure 4.7: Effective stress paths for varying B values and total stress ratios
(Folkes and Crooks 1985). .................................................................................... 4.11
Figure 4.8: Effective stress paths and pore water pressure generation associated
with different types of yield conditions (Leroueil and Tavenas 1986). ................ 4.11
Figure 4.9: Element of soil at depth z beneath on centreline of uniform circular
load (Muir Wood 1990) ........................................................................................ 4.13
Figure 4.10: Total stresses on centreline beneath uniform circular load on elastic
half space (Muir Wood 1990). .............................................................................. 4.14
Figure 4.11: Total (BCDE) and effective (B′C′D′E′) stress paths for a soil element
on the centreline beneath a circular load (Muir Wood 1990). .............................. 4.14
List of Figures Page xxiv

Figure 4.12: Variation of pore water pressure with applied surface load for
element on centreline beneath circular load (Muir Wood 1990). ......................... 4.15
Figure 4.13: Variation of the dimensional factor R (proportional to the horizontal
surface displacement at the toe) with factor of safety (Marche and Chapuis
1974) ..................................................................................................................... 4.17
Figure 4.14: Definition of the geometry and deformation parameters (Tavenas et
al 1979).................................................................................................................. 4.18
Figure 4.15: Average correlation between the maximum lateral movement at the
toe, ym, and settlement under the centre of the embankment, s, during
construction (Tavenas and Leroueil 1980)............................................................ 4.18
Figure 4.16: Effective stress and lateral displacement profiles at the end of
construction under (a) Cubzac-les-Ponts A and (b) Saint Alban B test fills
(Tavenas et al 1979) .............................................................................................. 4.19
Figure 4.17: Comparison of horizontal movement (on berm beyond embankment
toe) during loading and for 24 hour period for the Thames test embankment
(Marsland and Powell 1977) ................................................................................. 4.20
Figure 4.18: Observations of lateral displacement at the base of the embankment
with increasing embankment thickness at the Rio test embankment (Ramalho-
Ortigao et al 1983a)............................................................................................... 4.21
Figure 4.19: Contours of normalised shear-stress ratios at 4 m fill thickness from
finite element modelling at Muar-F test embankment, Malaysia (Indraratna et
al 1992).................................................................................................................. 4.22
Figure 4.20: Idealised section of embankment indicating the types of monitored
observations........................................................................................................... 4.23
Figure 4.21: (a) and (b) Excess pore water pressure under centre of embankment
versus applied embankment load. ......................................................................... 4.27
Figure 4.22: Settlement under the centre of the embankment versus relative
embankment height (H/Hf). ...................................................................................4.34
Figure 4.23: Vertical displacement at the embankment toe versus relative
embankment height. .............................................................................................. 4.35
Figure 4.24: Vertical displacement beyond the embankment toe versus relative
embankment height. .............................................................................................. 4.36
Figure 4.25: Lateral surface displacement at the embankment toe versus relative
embankment height. .............................................................................................. 4.38
Figure 4.26: Incremental vertical inclination angle (from inclinometer
observations at the embankment toe) with fill height (and time) for (a)
Thames and (b) King’s Lynn. ............................................................................... 4.39
List of Figures Page xxv

Figure 4.27: Vertical inclination angle on the surface of rupture from inclinometer
observations at the embankment toe versus relative embankment height. ........... 4.40
Figure 4.28: Maximum lateral displacement at the embankment toe versus
settlement under the centre of the embankment. ................................................... 4.43
Figure 4.29: Factor of safety versus relative embankment height ............................... 4.46
Figure 4.30: Failure model for post-failure deformation (Khalili et al 1996).............. 4.49
Figure 4.31: Effective stress paths post construction (Leroueil and Tavenas 1986).... 4.53
Figure 4.32: Excess pore water pressure response with time for several
piezometers at the Muar-EC test embankment. .................................................... 4.55
Figure 4.33: Excess pore water pressure response at Piezometer C1 with time at
the St. Alban-A test embankment. ........................................................................ 4.56
Figure 4.34: Schematic behaviour of strain rate effects on limit state surface with
decreasing strain rate in undrained creep tests (adapted from Leroueil and
Marques 1996) ...................................................................................................... 4.58
Figure 4.35: Relationship between the maximum lateral displacement at the toe,
ym, and settlement at the centre, s, at the end of construction (Tavenas et al
1979). .................................................................................................................... 4.60
Figure 4.36: Relationship between maximum lateral displacement (toe) and
settlement (centre) at the end of construction for different embankment
geometries and stability conditions (Tavenas et al 1979). .................................... 4.61

Figure 5.1: Comparison of population to slope stability incidents for earthfill,


rockfill and hydraulic fill embankments (excluding sloughing and unknown
embankment zonings). ............................................................................................ 5.6
Figure 5.2: Sloughing type failure (Foster 1999)........................................................... 5.7
Figure 5.3: Sliding block model showing kinetic energy lines for different
rheological models (Golder Assoc. 1995)............................................................. 5.11
Figure 5.4: Redistribution of the potential energy after failure (D’Elia et al 1998) .... 5.12
Figure 5.5: Contours of accumulated deviatoric plastic strain, ε D P , from finite
element analysis of an excavation: 3H to 1V slope, 10 m cut slope height, Ko
= 1.5, surface suction = 10 kPa (Potts et al 1997)................................................. 5.19
Figure 5.6: Distribution of water content near the slip place from failures in cut
slopes in London clay (Skempton 1964)............................................................... 5.20
Figure 5.7: Moisture content migration to the shear zone and decreasing shear
strength with time to failure from consolidated undrained tests on brown
London clay (Skempton and La Rochelle 1965)................................................... 5.20
List of Figures Page xxvi

Figure 5.8: Annual records of cut slope failures in heavily over-consolidated clays
in the UK (James 1970)......................................................................................... 5.21
Figure 5.9: Gradual reduction in the factor of safety with time due to soil
weakening and oscillations due to climatic effects (Picarelli et al 2000) ............. 5.22
Figure 5.10: Variation of the pore water pressure coefficient, ru, with time in
cuttings in Brown London clay (Skempton 1977). ............................................... 5.23
Figure 5.11: Time to first time failure versus cut slope angle for Type 1 and Type
2 slope failure geometries in London clay. ........................................................... 5.23
Figure 5.12: Idealised development of surface of rupture during construction for
embankments with wet placed fill layer/s............................................................. 5.26
Figure 5.13: Compound type surface of rupture for failures during construction at
(a) Scout Reservation dam (Mann and Snow 1992), and (b) Waco dam
(Stroman et al 1984).............................................................................................. 5.26
Figure 5.14: Carsington dam, section through the region of the initial failure at
chainage 725 m (Skempton and Vaughan 1993). ................................................. 5.27
Figure 5.15: Carsington dam; drained strength properties of intact “yellow clay”
and solifluction shears (Skempton and Vaughan 1993)........................................ 5.28
Figure 5.16: Contours of deviatoric strain predicted by finite element analysis at
chainage 725 m prior to the failure of Carsington dam (Potts et al 1990). ........... 5.29
Figure 5.17: Section through the failure in the upstream shoulder at San Luis dam
(courtesy of U.S. Bureau of Reclamation). ........................................................... 5.35
Figure 5.18: Drawdown causing failure and prior large drawdowns for (a) Dam
FD2, (b) Dam FD3, and (c) San Luis dam. ........................................................... 5.36
Figure 5.19: Composite shear strength envelope (Duncan et al 1990). ....................... 5.38
Figure 5.20: Effect of moisture migration on the undrained strength within the
shear band using idealised strength parameters. ................................................... 5.38
Figure 5.21: Type 1 failure slope geometries with different surface of rupture
orientations............................................................................................................ 5.45
Figure 5.22: Total post-failure deformation for slides in embankment dams during
construction. .......................................................................................................... 5.49
Figure 5.23: Maximum post failure velocity of slides in embankment dams during
construction. .......................................................................................................... 5.50
Figure 5.24: Residual drained strength from field case studies in clay soils in the
UK and ring shear tests on sand, kaolin and bentonite (Skempton 1985). ........... 5.51
Figure 5.25: Ring shear tests on normally consolidated sand-bentonite mixtures
(Skempton 1985, after Lupini et al 1981). ............................................................ 5.52
Figure 5.26: Section through Punchina cofferdam (adapted from Villegas 1982). ..... 5.54
List of Figures Page xxvii

Figure 5.27: Punchina cofferdam, (a) deformation and (b) velocity of the
downstream slope during construction (adapted from Villegas (1982))............... 5.54
Figure 5.28: Section through the failed region of Muirhead dam (Banks 1948) ......... 5.55
Figure 5.29: Muirhead dam, post failure deformation and velocity of the failure in
the upstream shoulder during construction. .......................................................... 5.55
Figure 5.30: Crest settlement monitoring of the slide at Waco dam (adapted from
Stroman et al 1984). .............................................................................................. 5.59
Figure 5.31: Carsington dam, deformation monitoring of upstream shoulder (a)
leading up to failure, (b) during failure (Skempton and Vaughan 1993), and
(c) the velocity of the slide mass........................................................................... 5.61
Figure 5.32: Section through failed region of Acu dam (Penman 1986) ..................... 5.62
Figure 5.33: Cross section of the slide during construction at Chingford dam
(Cooling and Golder 1942) ................................................................................... 5.63
Figure 5.34: Deformation and velocity of the slide at Lake Shelbyville dam
(adapted from Humphrey and Leonards (1986, 1988))......................................... 5.67
Figure 5.35: Post failure deformation and velocity of slide at Scout Reservation
dam (adapted from Mann and Snow (1992))........................................................ 5.67
Figure 5.36: Slides in embankment dams during drawdown; post failure (a)
deformation and (b) estimated velocity................................................................. 5.73
Figure 5.37: San Luis dam; the latter stages of post failure (a) deformation and (b)
velocity of the first time slide following the 1981 drawdown (data courtesy of
U.S. Bureau of Reclamation). ............................................................................... 5.76
Figure 5.38: Section through the slide at Dam FD2..................................................... 5.77
Figure 5.39: Deformation behaviour of the slide at Dam FD2 on reactivation
during the 1994 drawdown; (a) deformation and (b) velocity. ............................. 5.78
Figure 5.40: Upstream slope angle versus total post failure deformation for slides
in embankment dams during drawdown. .............................................................. 5.79
Figure 5.41: Post construction slides in the downstream shoulder; post failure (a)
deformation, and (b) estimated peak velocity. ......................................................5.84
Figure 5.42: Slide in the cut slope in London clay at Northolt in 1955 (Skempton
1964) ..................................................................................................................... 5.87
Figure 5.43: Post-failure slide deformation for cut slopes in London clays. ............... 5.90
Figure 5.44: Estimated peak post failure velocity of slides in cut slopes in London
clays....................................................................................................................... 5.90
Figure 5.45: Cut slope angle versus post-failure deformation at the slide toe for
cut slopes in heavily over-consolidated clays of Type 1 and Type 2 slope
failure geometry. ................................................................................................... 5.91
List of Figures Page xxviii

Figure 5.46: Cut slope angle versus estimated post-failure slide velocity for cut
slopes in heavily over-consolidated clays............................................................. 5.91
Figure 5.47: Failure in the cut slope in London clay at New Cross in 1841
(Skempton 1977)................................................................................................... 5.94
Figure 5.48: Failure in the retained cut slope in London clay at Wembley Hill
(Skempton 1977)................................................................................................... 5.95
Figure 5.49: Failure in the retained cut slope in London clay at Uxbridge in 1954
(adapted from Henkel 1956) ................................................................................. 5.97
Figure 5.50: Monitored post-failure deformation of the slide in the retained cut in
London clay at Uxbridge in 1954 (adapted from Watson 1956)........................... 5.97
Figure 5.51: Failure model for post-failure deformation (Khalili et al 1996).............. 5.98
Figure 5.52: Observed deformation versus the calculated residual factor of safety. .5.103
Figure 5.53: Schematic profile of slope failure showing travel distance and travel
distance angle ......................................................................................................5.106
Figure 5.54: H/L versus the slide volume for embankment dams and cut slopes of
moderate to rapid post-failure velocity. ..............................................................5.107
Figure 5.55: H/L versus the tangent of the fill or cut slope angle for embankment
dams and cut slopes of moderate to rapid post-failure velocity. .........................5.108

Figure 6.1: Determination of rockfill moduli, (a) modulus during construction,


Erc, and (b) modulus during reservoir filling, Erf (Fitzpatrick et al 1985)............... 6.5
Figure 6.2: Typical design of dumped rockfill CFRD (adapted from ICOLD
(1989b))................................................................................................................... 6.7
Figure 6.3: Typical zoning of main rockfill in current CFRD design practice for
construction with sound quarried rockfill (adapted from Cooke (1997); zoning
designators after Fell et al (1992)) .......................................................................... 6.8
Figure 6.4: Compression curves for dry and wet states, and collapse compression
from dry to wet state for Pyramid gravel in laboratory oedometer test (Nobari
and Duncan 1972a) ............................................................................................... 6.13
Figure 6.5: Settlement versus time curves for laboratory oedometer tests on
rockfill (Sowers et al 1965)................................................................................... 6.14
Figure 6.6: Typical stress-strain relationship of rockfill from a triaxial
compression test (Mori and Pinto 1988) ............................................................... 6.17
Figure 6.7: Finite element analysis of Foz do Areia CFRD (Saboya and Byrne
1993), (a) model and (b) stress paths during construction and first filling. .......... 6.18
Figure 6.8: Deformation modulus during construction (Erc) versus void ratio
(Pinto and Marques Filho 1998) ........................................................................... 6.20
List of Figures Page xxix

Figure 6.9: Ratio of deformation modulus on first filling to during construction


(Erf /Erc) versus valley shape factor (Pinto and Marques Filho 1998)...................6.20
Figure 6.10: Deformation modulus during construction (Erc) for Hydro Tasmania
CFRDs (Giudici et al 2000) .................................................................................. 6.21
Figure 6.11: Results of 3-dimensional finite element analysis of CFRD, to give
vertical displacement during construction versus valley shape (Giudici et al
2000) ..................................................................................................................... 6.22
Figure 6.12: Settlement rate analysis of Cedar Creek dam; (a) determination of
time to and (b) settlement rate versus time (Parkin 1977)..................................... 6.24
Figure 6.13: Post-construction crest settlement of membrane faced compacted
rockfill dams (Clements 1984).............................................................................. 6.26
Figure 6.14: Settlement index versus time for well-compacted rockfills (adapted
from Public Works Department NSW 1990) ........................................................ 6.27
Figure 6.15: Stress-strain relationship for rockfills observed during construction...... 6.33
Figure 6.16: Secant modulus versus vertical stress from monitoring during
construction........................................................................................................... 6.33
Figure 6.17: Tangent modulus versus vertical stress from monitoring during
construction........................................................................................................... 6.35
Figure 6.18: Idealised model for two-dimensional finite difference analysis of
cross-valley influence............................................................................................ 6.36
Figure 6.19: Results of two-dimensional finite difference analysis of the effect of
cross-valley shape on vertical stresses in the dam. (a), (c) and (e) represent
construction in 5 m lifts (to 100 m) and (b), (d) and (f) construction in a single
100 m lift. .............................................................................................................. 6.38
Figure 6.20: Indicators of cross valley arching effects from variations in the secant
modulus with vertical stress. ................................................................................. 6.39
Figure 6.21: Representative secant modulus (mostly Zone 3A rockfill) at end of
construction (Erc) versus D80 from average grading of the rockfill.......................6.42
Figure 6.22: Erf / Erc ratio versus embankment height ................................................. 6.45
Figure 6.23: Stress paths during construction and first filling for nominal 100 m
embankment with 1.3H to 1V upstream slope angle; (a) monitoring point
locations, (b) stress paths for points normal to face slab at 30% of the
embankment height. .............................................................................................. 6.49
Figure 6.24: Stress paths during construction and first filling for nominal 100 m
embankment with 1.55H to 1V upstream slope angle; (a) monitoring point
locations, (b) stress paths for points normal to face slab at 30% of the
embankment height. .............................................................................................. 6.49
List of Figures Page xxx

Figure 6.25: Face slab deflection during first filling (4/2/71 to 25/4/71) of Cethana
dam (Fitzpatrick et al 1973). ................................................................................. 6.50
Figure 6.26: Face slab deformation during first filling of Aguamilpa dam (Mori
1999) ..................................................................................................................... 6.51
Figure 6.27: Face slab deformation during first filling of Ita dam (Sobrinho et al
2000) ..................................................................................................................... 6.51
Figure 6.28: Scotts Peak dam; embankment zoning and location of face cracks
(courtesy of Hydro Tasmania)............................................................................... 6.52
Figure 6.29: Mackintosh dam (a) embankment section (courtesy of Hydro
Tasmania) and (b) face slab deformation on first filling (Knoop and Lack
1985) ..................................................................................................................... 6.52
Figure 6.30: Examples of derivation of zero time for post-construction settlement.... 6.56
Figure 6.31: Post-construction crest settlement versus time for dumped rockfill
CFRD, to at end of main rockfill construction. ..................................................... 6.57
Figure 6.32: (a) and (b) Post-construction crest settlement versus time for CFRD
constructed of compacted rockfills of medium to high intact strength, to at end
of main rockfill construction. ................................................................................ 6.58
Figure 6.33: (a) and (b) Post-construction crest settlement versus time for CFRDs
constructed of well-compacted rockfills of very high intact strength and of
well-compacted gravels, to at end of main rockfill construction. .......................... 6.59
Figure 6.34: Post-construction crest settlement versus time for dumped rockfill
CFRD, to at end of first filling. ..............................................................................6.60
Figure 6.35: Post-construction crest settlement versus time for CFRD constructed
of compacted rockfills of medium to high intact strength, to at end of first
filling. .................................................................................................................... 6.60
Figure 6.36: Post-construction crest settlement versus time for CFRDs constructed
of well-compacted rockfills of very high intact strength and of well-
compacted gravels, to at end of first filling. .......................................................... 6.61
Figure 6.37: Post construction crest settlement of Bastyan dam.................................. 6.65
Figure 6.38: Long-term crest settlement rate (as a percentage of embankment
height per log cycle of time) versus embankment height for compacted
rockfills ................................................................................................................. 6.66
Figure 6.39: Crest settlement attributable to first filling (excluding time dependent
effects)................................................................................................................... 6.68
List of Figures Page xxxi

Figure 7.1: Embankment construction indicating (a) broad layer width to


embankment depth ratio in the early stages of construction and (b) narrow
layer width to embankment depth ratio in the latter stages of construction.......... 7.24
Figure 7.2: (a) Total vertical stress (σz,EOC ) contours at end of construction for
1.8H to 1V embankment slopes, and (b) vertical stress ratio, Z1, versus h/H
under the embankment axis for various embankment slopes................................7.26
Figure 7.3: Embankment model for finite difference analysis during construction.....7.29
Figure 7.4: Total vertical stress, total lateral stress, and lateral (horizontal)
displacement distributions at end of construction for zoned earthfill finite
difference analysis................................................................................................. 7.30
Figure 7.5: Finite difference modelling results; (a) total vertical stress on the dam
centreline at end of construction, (b) settlement profile at dam axis at end of
construction, and (c) lateral deformation at Point a (Figure 7.3) during
construction. ..........................................................................................................7.32
Figure 7.6: Finite element analysis of zoned embankments highlighting arching of
core zone from (a) Beliche dam at about mid-height, elevation 22.5 m
(Naylor et al 1997), and (b) a puddle core embankment (Dounias et al 1996).....7.33
Figure 7.7: Estimated versus measured (from pressure cells) total vertical stresses
during construction................................................................................................ 7.34
Figure 7.8: Total vertical stress profiles from the numerical analysis of
construction for zoned embankments with different core slopes; (a) 0.2H to
1V case, (b) 0.5H to 1V case, and (c) total vertical stress at h/H = 0.70 with
increasing embankment height. ............................................................................. 7.36
Figure 7.9: Idealised types of pore water pressure response in the core during
construction. .......................................................................................................... 7.38
Figure 7.10: Pore water pressure response during construction in the core zone of
clayey earthfills placed at close to or wet of Standard optimum moisture
content. .................................................................................................................. 7.39
Figure 7.11: Triaxial isotropic compression tests with staged undrained loading
and partial drainage on a partially saturated silty sand (Bishop 1957). ................7.41
Figure 7.12: Estimated lateral displacement ratio of the core at end of
construction. .......................................................................................................... 7.45
Figure 7.13: Estimated lateral displacement ratio of the core versus fill height
above gauge during construction........................................................................... 7.45
Figure 7.14: Lateral displacement ratio and pore water pressure versus fill height
or measured total vertical stress at (a) Dartmouth dam; (b) Blowering dam;
(c) Talbingo dam and (d) Thomson dam............................................................... 7.50
List of Figures Page xxxii

Figure 7.15: Vertical stress versus strain of the core during construction for
selected cross-arm intervals of selected case studies; (a) dry placed clay
cores, and (b) dry placed dominantly sandy and gravelly cores with plastic
fines. ...................................................................................................................... 7.54
Figure 7.16: Confined secant moduli of the core versus vertical stress during
construction for selected cross-arm intervals of selected case studies, (a) dry
placed clay cores, and (b) dry placed dominantly sandy and gravelly cores
with plastic fines. .................................................................................................. 7.55
Figure 7.17: Seating settlements at low stresses suspected as cause of low moduli
estimates at low stresses........................................................................................ 7.56
Figure 7.18: Vertical strain versus effective vertical stress in the core at end of
construction for (a) dry placed clay earthfills and (b) dry placed dominantly
sandy and gravelly earthfills. ................................................................................ 7.58
Figure 7.19: Confined secant moduli versus effective vertical stress in the core at
end of construction for (a) dry placed clay earthfills and (b) dry placed
dominantly sandy and gravelly earthfills.............................................................. 7.59
Figure 7.20: Vertical strain in the core versus fill height during construction at
selected cross-arm intervals in selected case studies of earthfills placed close
to or wet of Standard optimum, for (a) thin core (combined slopes less than
0.5H to 1V), and (b) medium cores (combined slope ≥ 0.5H to 1V and < 1H
to 1V). ................................................................................................................... 7.62
Figure 7.21: Vertical strain in the core versus fill height at end of construction for
dominantly sandy and gravelly earthfills with plastic fines placed close to or
wet of Standard optimum, for (a) thin cores (combined slopes less than 0.5H
to 1V), and (b) medium to thick cores (combined slope ≥ 0.5H to 1V). ..............7.63
Figure 7.22: Vertical strain of the core versus fill height at end of construction for
clay earthfills placed close to or wet of Standard optimum moisture content,
for (a) thin core (combined slopes less than 0.5H to 1V), and (b) medium to
thick cores (combined slope ≥ 0.5H to 1V). ......................................................... 7.64
Figure 7.23: Core settlement during construction of earth and earth-rockfill
embankments (a) including Nurek dam, and (b) excluding Nurek dam. .............. 7.67
Figure 7.24: Effect of reservoir filling on a zoned embankment (Nobari and
Duncan 1972b) ...................................................................................................... 7.70
Figure 7.25: Water load acting on a homogeneous earthfill dam. ............................... 7.71
Figure 7.26: Lateral stress distribution at (a) end of construction, and (b) reservoir
full condition, for central core earth and rockfill embankment with core of
similar compressibility properties to the rockfill. ................................................. 7.72
List of Figures Page xxxiii

Figure 7.27: Lateral stress distribution at (a) end of construction, and (b) reservoir
full condition, for a central core earth and rockfill embankment with wet
placed core of low undrained strength. ................................................................. 7.73
Figure 7.28: Compression curves for dry and wet states and collapse compression
from the dry to wet state for Pyramid gravel in the laboratory oedometer test
(Nobari and Duncan 1972a). ................................................................................. 7.75
Figure 7.29: Idealised effect of matric suction on collapse compression of
earthfills (Khalili 2002)......................................................................................... 7.75
Figure 7.30: Cracking caused by post construction differential settlement between
the core and the dumped rockfill shoulders (Sherard et al 1963).......................... 7.77
Figure 7.31: Embankment model for finite difference analysis during first filling. .... 7.78
Figure 7.32: Model of the stress-strain relationship of the upstream rockfill
incorporating collapse compression (in one-dimensional vertical
compression). ........................................................................................................ 7.80
Figure 7.33: Case 8 – “dry” placed, very stiff core modelling collapse
compression in the upstream rockfill; numerical results of (a) vertical and (b)
lateral displacement on first filling........................................................................ 7.82
Figure 7.34: Case 9 – “dry” placed, very stiff core without collapse compression;
numerical results of (a) vertical and (b) lateral displacement on first filling........ 7.82
Figure 7.35: Case 10 – “wet” placed clay core of low undrained strength
modelling collapse compression in the upstream rockfill. Numerical results
of vertical and lateral deformation on first filling; (a) and (b) for reservoir at
40% of embankment height, (c) and (d) for reservoir at 98% of embankment
height. .................................................................................................................... 7.83
Figure 7.36: Case 11 – “wet” placed clay core without collapse compression;
numerical results of (a) vertical and (b) lateral displacement on first filling........7.84
Figure 7.37: Regional division of the embankment for analysis of post
construction surface deformation. .........................................................................7.85
Figure 7.38: Lateral displacement of the crest (centre or downstream region) on
first filling versus embankment height (displacement is after the end of
embankment construction). ................................................................................... 7.88
Figure 7.39: Lateral displacement of the downstream slope (mid to upper region)
on first filling versus embankment height (displacement is after the end of
embankment construction). ................................................................................... 7.95
Figure 7.40: Lateral displacement at the upper upstream slope of upstream crest
region on first filling versus embankment height (displacement after the end
of embankment construction)................................................................................ 7.97
List of Figures Page xxxiv

Figure 7.41: Lateral displacement of the crest (centre to downstream region) on


first filling versus downstream slope. ................................................................... 7.98
Figure 7.42: Lateral displacement of the downstream slope (mid to upper region)
on first filling versus downstream slope. .............................................................. 7.98
Figure 7.43: Post construction internal settlement of the core at Talbingo dam
(IVM ES1 at the main section)............................................................................7.101
Figure 7.44: Post construction internal settlement of the core at Copeton dam
(IVM B, 9 m downstream of dam axis at main section). ....................................7.101
Figure 7.45: Post construction internal settlement of the core at Bellfield dam. .......7.102
Figure 7.46: Post construction crest settlement at 3 years after end of construction,
(a) all data, (b) data excluding Ataturk. ..............................................................7.105
Figure 7.47: Post construction crest settlement at 10 years after end of
construction. ........................................................................................................7.105
Figure 7.48: Post construction crest settlement at 20 to 25 years after end of
construction. ........................................................................................................7.106
Figure 7.49: Post construction settlement of the downstream slope (mid to upper
region) at 3 years after end of construction. ........................................................7.107
Figure 7.50: Post construction settlement of the downstream slope (mid to upper
region) at 10 years after end of construction. ......................................................7.107
Figure 7.51: Post construction settlement of the upper upstream slope and
upstream crest region at 3 years after end of construction. .................................7.108
Figure 7.52: Post construction settlement of the upper upstream slope and
upstream crest region at 10 years after end of construction. ...............................7.108
Figure 7.53: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey earthfills; (a) all data,
(b) data excluding Ataturk. .................................................................................7.113
Figure 7.54: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey earthfills.......................7.114
Figure 7.55: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey sand to clayey gravel
(SC to GC) earthfills. ..........................................................................................7.114
Figure 7.56: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey sand to clayey gravel
(SC to GC) earthfills. ..........................................................................................7.115
Figure 7.57: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of silty sand to silty gravel (SM to GM)
earthfills...............................................................................................................7.116
List of Figures Page xxxv

Figure 7.58: Crest settlement versus time for zoned embankments with central
core zones of clayey earthfills of thick width (1 to 2.5H to 1V combined
width). .................................................................................................................7.116
Figure 7.59: Crest settlement versus time for zoned embankments with central
core zones of silty to clayey gravel and sand (SC, GC, SM, GM) earthfills of
thick width (1 to 2.5H to 1V combined width). ..................................................7.117
Figure 7.60: Crest settlement versus time for embankments with very broad core
widths (> 2.5H to 1V combined width) and limited foundation influence. ........7.118
Figure 7.61: Crest settlement versus time for embankments with very broad core
widths (> 2.5H to 1V combined width) and potentially significant foundation
influence..............................................................................................................7.118
Figure 7.62: Long-term post construction crest settlement rate for zoned earthfill
and earth-rockfill embankments of thin to thick core widths, (a) all data, and
(b) data excluding Ataturk. .................................................................................7.120
Figure 7.63: Long-term post construction crest settlement rate for embankments of
very broad core width, (a) all data, and (b) data excluding Belle Fourche,
Roxo and Rector Creek. ......................................................................................7.121
Figure 7.64: Crest settlement versus time for central core earth and rockfill
embankments with central core zones of clayey earthfills and steady post first
filling reservoir operation....................................................................................7.127
Figure 7.65: Crest settlement versus time for central core earth and rockfill
embankments with thin to thick central core zones of clayey earthfills
subjected to seasonal reservoir fluctuation..........................................................7.127
Figure 7.66: Post construction settlement and reservoir operation at Enders dam
(data courtesy of USBR). ....................................................................................7.131
Figure 7.67: Post construction settlement at main section and reservoir operation
at San Luis dam (data courtesy of USBR). .........................................................7.132
Figure 7.68: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey earthfills.......................7.133
Figure 7.69: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey earthfills.......................7.134
Figure 7.70: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey sand and clayey gravel
(SC/GC) earthfills. ..............................................................................................7.134
Figure 7.71: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey sand and clayey gravel
(SC/GC) earthfills. ..............................................................................................7.135
List of Figures Page xxxvi

Figure 7.72: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of silty sand and silty gravel (SM/GM)
earthfills...............................................................................................................7.135
Figure 7.73: Crest displacement versus time for zoned embankments with central
core zones of clayey earthfills of thick width (1 to 2.5H to 1V combined
width). .................................................................................................................7.136
Figure 7.74: Crest displacement versus time for zoned embankments with central
core zones of silty to clayey sand and gravel earthfills of thick width. ..............7.136
Figure 7.75: Crest displacement versus time for embankments with very broad
core zones (> 2.5H to 1V width) and with limited foundation influence............7.137
Figure 7.76: Crest displacement versus time for embankments with very broad
core zones (> 2.5H to 1V width) and with potentially significant foundation
influence on the embankment deformation. ........................................................7.137
Figure 7.77: Crest displacement normal to dam axis post first filling.......................7.139
Figure 7.78: Long-term settlement rates for the downstream slope (mid to upper
region) versus embankment height; (a) all data, and (b) data excluding Belle
Fourche................................................................................................................7.143
Figure 7.79: Post construction settlement of the downstream shoulder (mid to
upper region) for selected case studies................................................................7.144
Figure 7.80: Post construction displacement of the downstream shoulder (mid to
upper region) for selected case studies of embankments with very broad core
widths. .................................................................................................................7.145
Figure 7.81: Post construction displacement of the downstream shoulder (mid to
upper region) for selected case studies of central core earth and rockfill
embankments.......................................................................................................7.147
Figure 7.82: Post construction displacement of the downstream shoulder (mid to
upper region) for embankments with compacted earthfills and gravels in the
downstream shoulder. .........................................................................................7.148
Figure 7.83: Long-term settlement rates of the upper upstream slope to upstream
crest region of earthfill and earth-rockfill embankments. ...................................7.152
Figure 7.84: Post construction settlement of the upper upstream slope to upstream
crest region for embankments with poorly compacted rockfill in the upstream
slope. ...................................................................................................................7.153
Figure 7.85: Post construction settlement of the upper upstream slope to upstream
crest region for selected case studies (excluding poorly compacted rockfills)...7.153
Figure 7.86: Post construction lateral displacement of the upper upstream slope
and upstream crest region for selected case studies of embankments with
rockfill in the upstream shoulder.........................................................................7.156
List of Figures Page xxxvii

Figure 7.87: Post construction lateral displacement of the upper upstream slope to
upstream crest region for selected case studies of embankment with earthfills
and gravels in the upstream shoulder. .................................................................7.157
Figure 7.88: Selset Dam internal settlements during construction (data from
Bishop and Vaughan 1962).................................................................................7.159
Figure 7.89: “Idealised” model of collapse compression of poorly compacted
shoulder fill of puddle core earthfill dam on first filling.....................................7.164
Figure 7.90: “Idealised” model of yielding of poorly compacted shoulder fill and
puddle core on first drawdown of a puddle core earthfill dam. ..........................7.164
Figure 7.91: “Idealised” model of collapse compression on wetting of poorly
compacted earthfill in the downstream shoulder of a puddle core earthfill
dam. .....................................................................................................................7.165
Figure 7.92: Post construction crest settlement versus log time of puddle core
earthfill dams.......................................................................................................7.167
Figure 7.93: Definitions of “steady state” conditions during normal reservoir
operating conditions. ...........................................................................................7.172
Figure 7.94: Fluctuation of pore water pressures adjacent to and upstream of the
puddle core with fluctuations in reservoir level..................................................7.174
Figure 7.95: Ramsden Dam, internal horizontal deformation of the puddle core,
1988 to 1990 (Holton 1992). ...............................................................................7.176
Figure 7.96: Ramsden Dam, internal vertical strains measured in the puddle core
from 1988 to 1990 (data from Tedd et al 1997b and Kovacevic et al 1997). .....7.177
Figure 7.97: Vertical strain (or settlement) at the crest versus drawdown depth for
specific drawdown and refilling cycles (Tedd et al 1997b). ...............................7.179
Figure 7.98: Permanent vertical crest strains during cyclic drawdown (in normal
operating range) for puddle core dams with permeable upstream filling (after
Tedd et al 1997b). ...............................................................................................7.179
Figure 7.99: Ramsden Dam, settlement of SMPs versus time from 1988 to 1990
(data from Tedd et al 1990).................................................................................7.180
Figure 7.100: Ogden Dam, crest settlement versus time (Tedd et al 1997a). ............7.181
Figure 7.101: Idealised model of deformation during historically large drawdown..7.182
Figure 7.102: Walshaw Dean (Lower) dam, crest settlement versus time (Tedd et
al 1997a)..............................................................................................................7.183
Figure 7.103: Cross section through Beliche dam at the main section (Maranha
das Neves et al 1994) ..........................................................................................7.189
Figure 7.104: Cross section of Chicoasen dam at the main section (Moreno and
Alberro 1982)......................................................................................................7.189
List of Figures Page xxxviii

Figure 7.105: Section at Hirakud dam (Rao and Wadhwa 1958) ..............................7.192
Figure 7.106: Cumulative settlement during construction at Hirakud dam (adapted
from Rao 1957) ...................................................................................................7.192
Figure 7.107: Cross section of Nurek dam (adapted from Borovoi et al 1982).........7.193
Figure 7.108: Carsington dam, section at chainage 825 m after failure (adapted
from Skempton and Vaughan 1993). ..................................................................7.194
Figure 7.109: Carsington dam, vertical strain profiles from IVM during
construction, (a) IVM C at chainage 850 m, and (b) IVM B at chainage 705 m
(adapted from Rowe 1991)..................................................................................7.195
Figure 7.110: Carsington dam, variation of the maximum vertical strain in the
core with bank level (Rowe 1991). .....................................................................7.195
Figure 7.111: Maximum section at El Infiernillo dam (Marsal and Ramirez de
Arellano 1967). ...................................................................................................7.203
Figure 7.112: El Infiernillo dam, deformation instrumentation on station 0+135 at
the lower left abutment (Marsal and Ramirez de Arellano 1967).......................7.204
Figure 7.113: El Infiernillo dam, internal settlements during first filling at Station
0+135...................................................................................................................7.205
Figure 7.114: Main section at Chicoasen dam (Moreno and Alberro 1982)..............7.206
Figure 7.115: Chicoasen dam, inclinometer and cross-arm locations at the main
section (Moreno and Alberro 1982)....................................................................7.206
Figure 7.116: Internal settlement profiles during first filling at Chicoasen dam
(adapted from Moreno and Alberro 1982). .........................................................7.207
Figure 7.117: Beliche Dam, cross section (Maranha das Neves et al 1994)..............7.209
Figure 7.118: Beliche dam, post construction internal vertical settlement profile
within the embankment for the period from end of construction to 2.75 years
post construction. ................................................................................................7.209
Figure 7.119: Typical section of Canales dam (Bravo 1979). ...................................7.210
Figure 7.120: Canales dam, (a) cracking and differential settlement at the crest,
and (b) post construction settlement at the crest (Giron 1997) ...........................7.211
Figure 7.121: Main section at Djatiluhur dam (Sowers et al 1993)...........................7.212
Figure 7.122: Djatiluhur dam; (a) location of crack observed during construction,
January 1965; and (b) deformation of monuments at elevation 103 m, January
to April 1965 (Sherard 1973). .............................................................................7.214
Figure 7.123: Eppalock dam, post construction settlement of SMPs on the crest
and slopes for the first five years after construction. ..........................................7.215
Figure 7.124: Main section at Blowering dam (courtesy of NSW Department of
Public Works and Services, Dams and Civil Section). .......................................7.218
List of Figures Page xxxix

Figure 7.125: Blowering dam, vertical strain during construction for selected
cross-arms intervals in the core...........................................................................7.220
Figure 7.126: Blowering dam, post construction internal settlement of the core
from IVM A during first filling. ..........................................................................7.221
Figure 7.127: Cross section of Ataturk dam (Cetin et al 2000) .................................7.221
Figure 7.128: Post construction settlement of the crest and downstream shoulder
at Ataturk dam. ....................................................................................................7.223
Figure 7.129: LG-2 dam; crack in crest and possible shear plane in core (Paré
1984) ...................................................................................................................7.224
Figure 7.130: Main section at Copeton dam (courtesy of New South Wales
Department of Land and Water Conservation) ...................................................7.225
Figure 7.131: Copeton dam, post construction internal settlement profile in the
core at IVM A. ....................................................................................................7.227
Figure 7.132: Copeton dam, post construction settlement between cross-arms
versus time. .........................................................................................................7.228
Figure 7.133: Eppalock dam, original design at maximum section (Woodward
Clyde 1999).........................................................................................................7.230
Figure 7.134: Eppalock dam, post construction settlement at main section and
CS1......................................................................................................................7.231
Figure 7.135: Eppalock dam, post construction horizontal displacement at the
main section. .......................................................................................................7.233
Figure 7.136: Djatiluhur dam, post construction crest settlement..............................7.234
Figure 7.137: Bellfield dam, main section at chainage 701 m (courtesy of
Wimmera Mallee Water).....................................................................................7.235
Figure 7.138: El Infiernillo dam, internal (a) settlement and (b) displacement in
the core from 1966 to 1972 (Marsal and Ramirez de Arellano 1972) ................7.236
Figure 7.139: El Infiernillo dam, post construction (a) settlement and (b)
displacement of surface markers.........................................................................7.237
Figure 7.140: Main section at Eildon dam (courtesy of Goulburn Murray Water). ..7.239
Figure 7.141: Eildon dam, post construction settlement of SMPs at chainage 685
m..........................................................................................................................7.240
Figure 7.142: Eildon dam, internal settlement profiles in core from IVM records
for the period 1981 to 1998.................................................................................7.240
Figure 7.143: Main section at Svartevann dam (Kjœrnsli et al 1982). ......................7.243
Figure 7.144: Svartevann dam, post construction (a) settlement and (b)
displacement. .......................................................................................................7.246
Figure 7.145: Cross section at Cougar dam (Cooke and Strassburger 1988). ...........7.247
List of Figures Page xl

Figure 7.146: Cougar dam; post construction (a) settlement and (b) displacement
normal to dam axis of the crest at the maximum section (adapted from Pope
1967). ..................................................................................................................7.248
Figure 7.147: Cougar dam; differential settlement and cracks on crest (Pope 1967) 7.249
Figure 7.148: Wyangala dam, post construction internal settlement in IVM B
located 9 m upstream of dam axis.......................................................................7.250
Figure 7.149: Section at Wyangala dam (courtesy of New South Wales
Department of Land and Water Conservation). ..................................................7.251
Figure 7.150: Rector Creek dam main section (Sherard 1953)..................................7.254
Figure 7.151: Rector Creek dam; post construction settlement and displacement
versus log time (adapted from Sherard et al 1963). ............................................7.254
Figure 7.152: Roxo dam, settlement versus log time (adapted from De Melo and
Direito 1982). ......................................................................................................7.255
Figure 7.153: Main section at Mita Hills dam (Legge 1970).....................................7.256
Figure 7.154: Main section at Dixon Canyon dam (courtesy of U.S. Bureau of
Reclamation). ......................................................................................................7.258
Figure 7.155: Post construction (a) settlement and (b) displacement of SMPs on
the upper upstream slope at the Horsetooth Reservoir embankments. ...............7.259
Figure 7.156: Belle Fourche dam; main section as constructed in 1911 (courtesy
of U.S. Bureau of Reclamation)..........................................................................7.260
Figure 7.157: Belle Fourche dam; post construction crest settlement over the
period 1911 to 1928 (to 17 years post construction)...........................................7.264
Figure 7.158: Belle Fourche dam closure section; post construction (a) settlement
and (b) displacement normal to dam axis of the crest and slopes over the
period 1985 to 2000 (73 to 89 years post construction)......................................7.265
Figure 7.159: San Luis dam, section of the slide in the upstream slope in 1981 at
Station 135+00 (courtesy of United States Bureau of Reclamation). .................7.266
Figure 7.160: San Luis dam, difference between actual and predicted settlement
(Von Thun 1988).................................................................................................7.268
Figure 7.161: San Luis dam, comparison of settlement between SMPs at 13 m
upstream of dam axis and SMPs at 6.5 m downstream. ......................................7.268
Figure 7.162: San Luis dam, post construction settlement of SMPs in the vicinity
of the slide area. ..................................................................................................7.269
Figure 7.163: Main section at Horsetooth dam (courtesy of the United States
Bureau of Reclamation). .....................................................................................7.271
Figure 7.164: Horsetooth dam, post construction displacement of SMPs near to
the main section...................................................................................................7.272
List of Figures Page xli

Figure 7.165: Pueblo dam; cross section of left abutment embankment at Station
75.........................................................................................................................7.273
Figure 7.166: Pueblo dam, post construction settlement of the left embankment
near Station 75.....................................................................................................7.274
Figure 7.167: Yan Yean dam; settlement rate versus time of SMPs at Chainage
150 and 750 m. ....................................................................................................7.279
Chapter 1: Introduction Page i

TABLE OF CONTENTS
1.0 INTRODUCTION ....................................................................... 1.1
1.1 OBJECTIVES OF THE R ESEARCH ....................................................................... 1.1
1.1.1 Landslides in Soil Slopes ........................................................................... 1.2
1.1.2 Deformation Behaviour of Embankment Dams......................................... 1.3
1.2 STRUCTURE OF THE THESIS ............................................................................. 1.4
1.3 TERMS AND DEFINITIONS ................................................................................ 1.5
1.3.1 General Terms and Definitions ................................................................. 1.5
1.3.2 Terms and Definitions for Landslides ....................................................... 1.5
1.3.3 Terms and Definitions for Embankments .................................................. 1.9

LIST OF TABLES
Table 1.1: IUGS (1995) velocity classifications for landslides. .................................... 1.6
Table 1.2: Classification of unconfined compressive strength of intact rock (AS
1726-1993) ............................................................................................................ 1.13

LIST OF FIGURES
Figure 1.1: Landslide classification system and main slide types.................................. 1.7
Figure 1.2: Definition of travel distance, travel distance angle, and slope
geometry. ................................................................................................................. 1.8
Figure 1.3: Slope failure geometries (a) Type 1 - failure at the top of the cut or fill
slope, (b) Type 2 – the toe of failure is coincident with the toe of the slope,
and (c) Type 5 – the surface of rupture extends below and daylights beyond
the toe of the slope. ................................................................................................. 1.8
Figure 1.4: Classification of slope failure geometries (a) Type 3 –travel on
relatively uniform or gradually decreasing slope angles, and (b) Type 4 –
travel on changing slope angles from steep to shallow...........................................1.9
Figure 1.5: Embankment zoning classification system (Foster 1999) ......................... 1.10
Figure 1.6: Dam zoning categories of embankment types (Foster et al 2000).............1.10
Chapter 1: Introduction Page 1.1

1.0 INTRODUCTION

1.1 OBJECTIVES OF THE R ESEARCH


This thesis is a study of the pre and post failure deformation behaviour of landslides in
cut, fill and natural soil slopes, and of the deformation behaviour of embankment dams.
The need for the research on the deformation behaviour of landslides stems from
recognition of the deficiency of quantitative information databases for risk assessment
methods in evaluation of landsliding in soil slopes. The fundamental objective of the
research was to address these deficiencies for a number of slope conditions and material
types. No risk assessment of itself has been undertaken.
The main objectives of the research on the deformation behaviour of landslides in
soil slopes were, for selected classes of slope and soil characteristics, to:
i) Develop methods or guidelines for assessment/evaluation of the pre-failure
deformation behaviour leading up to landsliding;
ii) Identify the slope conditions and soil characteristics that will result in “slow” or
“rapid” landsliding; and
iii) Develop methods for prediction of the post failure deformation behaviour.

For embankment dams, those who are responsible for interpreting the deformation
behaviour do not have clear criteria to assess their data and often have to rely on their
judgement and personal experience. Whilst numerical modelling methods make for
useful tools in modelling embankment dam behaviour, few reported Type A predictions
have been able to accurately model the deformation behaviour during construction and
on first filling. Methods based on comparison to historical records of similar
embankments are still heavily relied upon and, particularly for the assessment of post
construction deformation behaviour, provide the responsible personnel with the best
guide to assessment of embankment performance.
The main objectives of the research on the deformation behaviour of embankment
dams were, from the historical performance records of case studies of selected
embankment types, to:
i) Broadly define “normal” deformation behaviour and from this platform to then
identify “abnormal” deformation behaviour;
Chapter 1: Introduction Page 1.2

ii) Define trends in the deformation behaviour that are potentially indicative of a
marginally stable or unstable slope condition, and are precursors to potential slope
instability; and
iii) Provide owners and their consultants with methods that allow comparison of the
deformation behaviour of their structures to similar embankment types during and
post construction.

The method of research was case study driven. Outcomes were developed from
analysis of the database in consideration of the mechanism/s influencing the
deformation behaviour. The database of landslides included some 450 case studies and
of embankment dams some 170 case studies.

1.1.1 LANDSLIDES IN SOIL SLOPES

For the landslides in constructed and natural soil slopes, information for each case study
was collected on:
• Slope geometry, both in the source area and the downslope travel path;
• Geology;
• Hydrogeology;
• Geotechnical properties of the soils within which the landslide occurred, and on the
downslope path for slides that travelled large distances beyond the toe of the slide
area;
• Slope history;
• Characteristics of the surface of rupture;
• Mechanism and trigger of the slope failure, and mechanism of the post failure travel;
and
• Deformation behaviour pre, at and post failure, both actual records and anecdotal
information.

Most of the case study data was from the published literature, but for a minority of
cases it was obtained or enhanced from unpublished reports from sponsors or in-kind
sponsors of the project.
Monitored records of pre-failure deformation were available for a limited number of
cases only. These were generally from well-instrumented cut or fill slopes constructed
Chapter 1: Introduction Page 1.3

to failure, but also from several of the failures in embankment dams during construction
and failures in active waste spoil piles. Anecdotal information from eyewitness
accounts was also available for several of the case studies.
For the post failure deformation behaviour of landslides case studies with actual
monitoring records were also limited. In most cases the data is limited to the travel
distance and areal extent of the failed slide mass. Some data is available on the actual
slide velocity from measurement points on “slow” moving slides, but is mostly from
estimates based on eyewitness accounts or the super-elevation on bends in the travel
path of fast moving confined debris flows or flow slides. Delineation in the travel
velocity of the slide mass is therefore often as either “slow” or “rapid”.
The slope classes, soil and slide types considered in the research were for landslides
in:
• Soil types that are potentially contractile on shearing under the stress conditions
imposed, including loose placed granular fills, mine waste stockpiles, hydraulic fills,
submarine slopes and sensitive clays.
• Cut slopes in residual soils, colluvium and weathered rock;
• Cut slopes in high plasticity, heavily over-consolidated clays;
• “Slides of debris”;
• Fill embankments constructed on soft ground; and
• Embankment dams, including failures during construction, on drawdown and post
construction in the downstream shoulder.

For several of the slope classes (e.g. landslides in sensitive clays) the current
literature on deformation behaviour has been well researched. Inclusion of these slide
types is mainly for comparative purposes to other slide groups.

1.1.2 DEFORMATION BEHAVIOUR OF EMBANKMENT DAMS

The analysis of embankment dams was on the following embankment types:


• Earthfill, zoned earthfill and zoned earth and rockfill embankments of rolled
earthfill construction;
• Puddle core earthfill embankments;
• Concrete face rockfill dams.
Chapter 1: Introduction Page 1.4

The deformation behaviour during construction, on first filling and post first filling
was analysed for the above embankment types.
About 40 to 50% of the case studies were from Australian dam authorities / owners
and the United States Bureau of Reclamation for which detailed information was made
available on material types, construction procedures and instrumentation records. This
was supplemented with case studies from the published literature for which the
available information was of variable quality.

1.2 STRUCTURE OF THE THESIS


The thesis structure is as follows:
• A brief review of the literature pertaining to the broader concepts of landslides in
soil slopes and deformation behaviour of embankment dams (Chapter 2). A more
detailed literature review is presented at the beginning of each of Chapters 3 through
7.
• Landslides in soil slopes of “rapid” post failure velocity (Chapter 3). From analysis
of the database, review of the literature and in consideration of the mechanics of
failure, guidelines are developed for slope geometrical conditions, soil
characteristics and the pre-failure deformation behaviour for the classes of “rapid”
landslides considered. Methods are developed for estimation of the post failure
travel distance that incorporate the quantifiable factors thought to significantly affect
the post failure travel for a specific slide class.
• Failures in fill embankments constructed on soft ground (Chapter 4). From analysis
of the pre-failure deformation behaviour guidelines are developed for identification
of an impending failure condition. Guidelines are presented in Chapter 5 for
prediction of the post failure deformation.
• Landslides in cut slopes of heavily over-consolidated high plasticity clays, and
failures in embankment dams (Chapter 5). The analysis concentrates on the
mechanics of failure and post failure deformation behaviour. Guidelines are
presented for prediction of the post failure deformation.
• The deformation behaviour of rockfill (Chapter 6). Guidelines are developed for
estimation of the moduli of rockfill during construction and prediction of
deformation behaviour during and post construction developed from the
deformation records of mostly concrete face rockfill dams.
Chapter 1: Introduction Page 1.5

• The deformation behaviour of embankment dams (Chapter 7). From a large


database of case studies, methods are developed and existing methods enhanced for
evaluation and prediction of the deformation behaviour of an embankment during
and post construction. Guidelines are presented for identification of potentially
“abnormal” deformation behaviour, and the potential indicators of a marginally
stable slope condition.

Each of Chapters 3 through 7 were published as individual reports through the


School of Civil and Environmental Engineering at The University of New South Wales
(Hunter et al (2000); Hunter and Fell (2001, 2002a, 2002b, 2003)).

1.3 TERMS AND DEFINITIONS

1.3.1 GENERAL TERMS AND DEFINITIONS

The classification of soils is in accordance with the Australian soil classification system
(Australian Standard AS 1726-1993 - Geotechnical Site Investigations). This
classification system is similar to the Unified Soil Classification System (USBR Earth
Manual and ASTM D2487-69) except that the particle size limits for the sand and
gravel sizes are in metric units and at slightly different sizes.

1.3.2 TERMS AND DEFINITIONS FOR LANDSLIDES

Following is a summary of definitions of terms. Details are given in Appendix A.


Landslide Velocity. The post failure velocity of the slide mass is classified
according to the system proposed by the International Union of Geological Sciences
(IUGS 1995), as presented in Table 1.1.
“Rapid Landslide” – a landslide that has a post failure travel velocity in the rapid,
very rapid or extremely rapid class as defined by IUGS (1995). Most of what are
termed “rapid” landslides would classify as very rapid or extremely rapid, having post
failure velocities greater than 3 metres/minute and generally in the order of metres/sec.
“Slow Landslide” – a landslide that has a post failure travel velocity in the
moderate, slow or very slow class as defined by IUGS (1995), in the range from
centimetres/year to less than 1.8 metres/hour.
Chapter 1: Introduction Page 1.6

Table 1.1: IUGS (1995) velocity classifications for landslides.


Velocity Description of Velocity limits Value in mm/sec
Classification Velocity
7 Extremely rapid > 5 m/sec > 5x10 3
6 Very rapid 3 m/min to 5 m/sec 50 to 5x103
5 Rapid 1.8 m/hour to 3 m/min 0.5 to 50
4 Moderate 13 m/month to 1.8 m/hour 5x10-3 to 0.5
3 Slow 1.6 m/year to 13 m/month 50x10-6 to 5x10-3
2 Very slow 16 mm/year to 1.6 m/year 0.5x10-6 to 50x10-6
1 Extremely slow ≤ 16 mm/year ≤ 0.5x10-6

Landslide Classification is according to Hutchinson (1988). For “rapid” landslides a


dual classification is used to describe the initiating landslide in the source area and the
subsequent slide movement in the travel region. The main terms used to describe the
movement of a landslide are shown in Figure 1.1 and described below.
Dilative and Contractile Slides – description of the initial tendency of the soil on the
surface of rupture to increase (dilate) or decrease (contract) in volume under drained
shear, or to develop negative or positive pore water pressures in undrained shear when
in a saturated (or near saturated) condition.
Flow Slide is used to describe the initiating slide in saturated (or near saturated)
contractile soils, where the failure or rapid acceleration of the slide mass occurs as a
result of a large loss in undrained strength due to static liquefaction on shearing. Flow
slide is also used as a travel classification descriptor to describe landslides where the
main volume of the slide mass is bodily carried on a liquefied basal zone, but is only
applicable to landslides initially classified as flow slides (Figure 1.1).
Defect Controlled Slide is used to describe landslides that are dominantly controlled
by defects in the soil or weathered rock mass.
“Slide of Debris” is a general term that is used to describe landslides in colluvium,
talus or other slope mantling debris.
Debris Flow is used to describe turbulent post failure slide movements of a
combination of water, air and debris. The slide mass is a broken up mass of material
that no longer retains its original structure or fabric.
Chapter 1: Introduction Page 1.7

Debris Slide is used to describe sliding on a defined basal surface where the slide
mass retains its structure and fabric during travel. Debris Slide is generally used as a
descriptor for slides that travel beyond the source area, and intact slide for slides where
the post failure deformation is largely confined to the source area.

Figure 1.1: Landslide classification system and main slide types.

Travel distance, L, landslide height, H, and travel distance angle, a, defined in Figure
1.2, are used to describe the overall longitudinal geometry of the slide mass from the
crest of the source area to the distal toe of the travel.
Source area slope angle (a 1 ), rupture surface inclination (a base ), initial downslope
angle (a 2 ) and distal downslope angle (a 3 ), defined in Figure 1.2, and cut (a cut ) or fill
slope angle (a fill ), are used to describe the longitudinal slope geometry. Note that a 3 is
undefined for several failure slope geometries (see below).
Slope failure geometry or failure type is a classification of the position of the initial
failure within the slope and the longitudinal shape of the downslope geometry. Five
types have been defined. Types 1, 2 and 5 are defined in Figure 1.3, and Types 3 and 4
in Figure 1.4. For Types 1 and 2 the slope below the cut of fill slope is near horizontal.
For “rapid” landslides with relatively long travel paths Types 1, 2 and 4 are
characterised by sharp changes in slope angle whilst Type 3 is characterised by
relatively smooth changes. Further details are given in Appendix A.
Chapter 1: Introduction Page 1.8

Figure 1.2: Definition of travel distance, travel distance angle, and slope geometry.

Figure 1.3: Slope failure geometries (a) Type 1 - failure at the top of the cut or fill slope,
(b) Type 2 – the toe of failure is coincident with the toe of the slope, and (c) Type 5 –
the surface of rupture extends below and daylights beyond the toe of the slope.
Chapter 1: Introduction Page 1.9

Figure 1.4: Classification of slope failure geometries (a) Type 3 –travel on relatively
uniform or gradually decreasing slope angles, and (b) Type 4 – travel on changing slope
angles from steep to shallow.

Degree of confinement of the travel path - the terms used (after Golder Assoc. 1995)
are; confined – the travel path is constrained by the relatively steep side-slopes of a
gully or small valley; unconfined – the travel path is on open slopes; and partly confined
– the travel path is not sharply defined by a topographic depression. Further details are
given in Appendix A.

1.3.3 TERMS AND DEFINITIONS FOR EMBANKMENTS

(a) Embankment Dam Zoning Classification


The embankment dam zoning classification system used is essentially the system
developed by Foster (1999). The system incorporates three components to describe the
general zoning of the embankment; embankment type, embankment filters and
Chapter 1: Introduction Page 1.10

foundation filters, and consists of a three number system (Figure 1.5). The codes to the
numbering used for the embankment zoning category are given in Figure 1.6.

Figure 1.5: Embankment zoning classification system (Foster 1999; Foster et al 2000)

core
downstream zone
of rockfill Puddle core

0. Homogeneous earthfill 4. Zoned earth and rockfill 8. Puddle core earthfill

Embankment filter rockfill concrete corewall


and/or rockfill
Foundation filter earthfill
core

1. Earthfill with filter 5. Central core earth and rockfill 9. Earthfill with corewall

concrete concrete corewall


facing
Rock toe rockfill
earthfill
Max 0.2H

2. Earthfill with rock toe 6. Concrete face earthfill 10. Rockfill with corewall

concrete
core facing hydraulic fill core
downstream zone
of sand/gravel
rockfill

3. Zoned earthfill 7. Concrete face rockfill 11. Hydraulic fill

Figure 1.6: Dam zoning categories of embankment types (Foster 1999)

For the filter classification the numbering code system used for both the embankment
and foundation filters is as follows:
0 - No filter drains present
1 - One filter present
Chapter 1: Introduction Page 1.11

2 - Two (or more) filter zones present


X - Unknown

A sub-classification system has been developed to describe the thickness or width of


the core. It was developed mainly to distinguish between zoned earthfill or earth and
rockfill embankments with thin central cores from those with thick central cores. It was
then broadened to encompass all earthfill and zoned earth-rockfill embankments. The
core width classification is:
• Thin earthfill core – symbol c-tn. Core width increasing at less than or equal to 0.5
times the vertical distance below the crest; i.e. slopes less than or equal to 0.25H to
1V for central cores (H to V = horizontal to vertical).
• Medium thickness earth core – symbol c-tm. Core width increase in the range 0.5 to
1.0 times the vertical distance below the crest. For central cores it includes cores
with slopes greater than 0.25H to 1V and less than or equal to 0.5 H to 1V.
• Thick earthfill core – symbol c-tk. Core width increase in the range 1.0 to 2.5 times
the vertical distance below the crest. For central cores it includes cores with slopes
greater than 0.5H to 1V and less than 1.25 H to 1V.
• Very broad earthfill core - symbol c-vb. This classification includes all
homogeneous earthfill embankments and earthfill embankments with filters as well
as zoned earthfill and zoned earth and rockfill embankments where the main
earthfill water barrier zone has an average width equal to or greater than the width at
the crest plus 2.5 times the depth below the crest.

For zoned earthfill embankments the core width classification only considers the
width of the main water barrier zone, from its upstream edge to the downstream edge of
the filter or permeable transition zones. Therefore, zoned embankments with filters
separating the core from similar type earthfills in the downstream shoulder may not be
classified as “very broad”, although, earthfill embankments with chimney filters are
classified as “very broad”. This is somewhat contradictory but the database only
includes one earthfill embankment with a chimney filter (Mita Hills) classified as “very
broad” that would fail the zoned embankment criteria for the “very broad”
classification.
Chapter 1: Introduction Page 1.12

(b) Classification of Rockfill Placement


The method of placement of rockfill has a significant influence on its compressibility
during construction and its deformation behaviour post construction. The definitions by
Cooke (1993, 1984) for dumped and compacted rockfill have been used as a basis for
categorisation of the method of placement. The definitions used are:
• Compacted Rockfill – rockfill placed in layers up to 2 m thickness (generally 0.9 to
2.0 m thick) and compacted by smooth drum vibrating roller (SDVR). Accepted
practice is typically 4 to 6 passes of a minimum 10 tonne (possibly up to 15 tonne)
deadweight vibrating roller, with variation in layer thickness, added water and
number of passes depending on the quality and type of the rockfill, amount of fines
and location within the embankment. Three classifications for compacted rockfill
have been used:
− Well-compacted – layer thickness typically less than about 1.0 m (depending on
the compressive strength of the intact rock) and compacted with a minimum of
four passes of a 10 to 15 tonne deadweight SDVR.
− Reasonable Compaction – layer thickness typically 1.5 to 2.0 m and compacted
with typically four passes of a 10 tonne SDVR.
− Reasonably to Well Compacted - layer thickness typically 1.2 to 1.6 m
(depending on the compressive strength of the intact rock) and compacted with
typically 4 to 6 passes of a 10 to 15 tonne SDVR.
• Rockfill not formally compacted or “poorly compacted”. Several methods of
rockfill placement have been included under the definition “poorly compacted”,
these include:
− Dumped rockfill – rockfill placed in lifts ranging from several to tens of metres
thickness, with or without sluicing, and without formal compaction.
− Rockfill placed in lifts less than about 2 to 3 m thickness and not formally
compacted (i.e. without the use of rollers for compaction). Specified track
rolling by bulldozer or other plant, or rockfill indicated as being trafficking by
trucks or other haulage equipment has been classified under “not formally
compacted”.
− Rockfill placed in lifts greater than 2 to 3 m and formally compacted. For these
rockfills the lift thickness is considered too great for compaction to have any
significant influence at depth.
Chapter 1: Introduction Page 1.13

Watering is an important component for placement of rockfills, particularly in cases


where the compressive strength of the rock used in the rockfill is susceptible to
reduction on wetting, breaks down under the action of the roller, or if the rockfill
contains large quantities of fines. Cooke (1993) comments that watering is not overly
important for compaction of very high strength rockfills that are not susceptible to
weakening on wetting. However, these rockfills can still show collapse type settlements
on wetting.
For dumped rockfills, sluicing had a significant influence on the deformation
behaviour of the rockfill as evidenced by the large collapse deformations of dry dumped
or poorly sluiced rockfills when wetted (Cogswell dam (Baumann 1958), Strawberry
and Dix River dams (Howson 1939)).

(c) Unconfined Compressive Strength of Rock


The classification system from Australian Standard AS 1726-1993 is used to define the
unconfined compressive strength (UCS) of intact rock used in rockfill. The descriptors
and the UCS range they represent are given in Table 1.2.

Table 1.2: Classification of unconfined compressive strength of intact rock (AS 1726-
1993)
Strength Descriptor UCS Range (MPa)
Extremely High > 240
Very High 70 to 240
High 20 to 70
Medium 6 to 20
Chapter 2: Literature Review Page i

TABLE OF CONTENTS
2.0 LITERATURE REVIEW ........................................................... 2.1
2.1 LANDSLIDES IN SOIL S LOPES ........................................................................... 2.1
2.2 DEFORMATION BEHAVIOUR OF EMBANKMENT DAMS ..................................... 2.7
2.3 WHERE T HIS R ESEARCH F ITS IN ......................................................................2.9

LIST OF FIGURES
Figure 2.1: Schematic of the system for geotechnical characterisation of slope
movement (Leroueil et al 1996).............................................................................. 2.3
Figure 2.2: The stages of slope movement (Leroueil et al 1996)................................... 2.4
Figure 2.3: Concept of the creep model under constant deviatoric stress
conditions. ............................................................................................................... 2.6
Figure 2.4: Monitored deformation behaviour at retained cut slope failure in
London clay at Kensal Green (Skempton 1964) ..................................................... 2.7
Chapter 2: Literature Review Page 2.1

2.0 LITERATURE REVIEW


The following literature review is a summary of the broader concepts related to the
deformation behaviour of landslides and embankment dams. A more detailed review of
the literature oriented toward the chapter subject is presented at the beginning of each of
Chapters 3 through to 7.

2.1 LANDSLIDES IN SOIL SLOPES


The slope movement classification systems of Varnes (1978) and Hutchinson (1988) are
primarily based on geomorphological classification of the type of slope movement with
lesser considerations of the mechanism/s of slope movement, material types, stages of
movement and rates of movement. The system of Varnes (1978) characterises the type
of movement into falls, topples, slides, spreads and flows, and material types into
bedrock, debris (coarse grained soils) and earth (fine grained soils). Hutchinson (1988)
expanded on aspects of the Varnes system to incorporate information on aspects of the
post failure behaviour of several classes of landslides including the relationship between
sediment concentration and flow type, material gradation, rate of movement and post
failure travel distance.
With the gradual shift toward risk based methods of assessment of landslides it
became evident that the geomorphologically based classification systems available at
the time, which are fundamental to any slope movement classification system, were not
sufficient in their current form. Leroueil (2001) comments that he and his colleagues
started working toward an expert system on slope stability in the late 1980’s, the
outcome of which is the geotechnical characterisation system described by Leroueil et al
(1996) (after Vaunet et al. 1994). Before briefly commenting on the geotechnical
characterisation of slope movement system, the comments by Leroueil (2001) in
relation to the problem at hand and the state of the published literature at the time (in the
mid to late 1980’s) are interesting; he states that:

“… we rapidly realised that the problem involves a large variety of


geomaterials and many types of slope movement under a variety of climatic
conditions, that the problem is complex and controlled by laws and
parameters that vary with the stage of movement, and the type of
Chapter 2: Literature Review Page 2.2

geomaterial, that the relevant information was extremely scattered in the


literature, and that solutions to problems related to slopes have often been
developed on a local or regional basis.”

Another important point to acknowledge that has been raised in a number of keynote
papers / lectures is the need for integration between geo-disciplines in the study of
landslides. Fell et al (2000), when commenting on the priority areas of research and
development associated with landsliding, emphasise the “overwhelming need to
integrate geology, geomorphology and geotechnical engineering in these studies”.
Leroueil (2001) in his Rankine lecture states:

“… slope movements as such are mechanical responses of soil or rock to


changes in geometry, boundary conditions, pore pressures or strength
parameters with time. They have thus also to be examined from a
mechanical or geotechnical viewpoint. In fact, good understanding of slope
movements can be obtained only from a joint effort from geologists,
geomorphologists and geotechnical engineers. As water is a major factor in
slope behaviour, the contribution of hydrologists and hydrogeologists is
also important. Slope movements must thus be seen as a multidisciplinary
domain in which the role of the geotechnical engineer is to improve our
understanding of soil behaviour in this specific context and to try and
minimize the economic and social impacts of slope movements.”

The system for geotechnical characterisation of slope movement as described by


Leroueil et al (1996) is the integration of movement types, material types and movement
stages into a 3-dimensional matrix, as shown in Figure 2.1. The movement stages
(Figure 2.2) are:
• Pre-failure – representing the slope deformation behaviour for first time slides prior
to the slope failure;
• At failure or the onset of failure – “characterised by the formation of a continuous
shear band or surface through the entire soil mass” (Leroueil et al 1996);
• Post failure – representing the movement of the slide mass from the failure
condition until it essentially stops; and
• Reactivation – renewal of movement of a quiescent landslide mass along one or
more pre-existing shear surfaces. Picarelli (2000) comments that this stage may
include a pre and post failure stage of movement.
Chapter 2: Literature Review Page 2.3

The earlier described geomorphological classification systems form the movement


type categories. Leroueil et al (1996) considered it necessary to expand the material
type categories used in the earlier systems to “take into account the mechanical
characteristics of soils and rocks” known to influence the deformation behaviour under
the boundary conditions imposed (e.g. the influence of over-consolidation, soil structure
and discontinuities on the deformation behaviour of natural and cut slopes in clays).
For each “active” element within the matrix there is a characterisation sheet (Figure
2.1) that contains information on the factors affecting or controlling the slope
movement, the potential evidentiary signs of movement and the possible consequences
of the movement. The basis of the characterisation for each element is an understanding
of the mechanisms involved and the mechanics of the deformation.
From a risk assessment point of view, the characterisation system provides a
framework within which the components (e.g. the hazard, its probability of failure and
the consequences of that failure) can be analysed and evaluated. It allows for informed
judgements to be made related to the potential for and consequences of landsliding
including the type and location of monitoring instrumentation, warning systems and
possible slope remedial measures. It also identifies deficiencies in the current state of
knowledge that can be targeted for research investment.

Figure 2.1: Schematic of the system for geotechnical characterisation of slope


movement (Leroueil et al 1996)
Chapter 2: Literature Review Page 2.4

Figure 2.2: The stages of slope movement (Leroueil et al 1996)

The outcomes of the risk assessment framework of analysing landsliding is evident in


the structure of a number of keynote papers of recent years, for example Fell et al
(2000), Picarelli (2000) and Leroueil (2001). These papers are aimed at bringing
together the knowledge published in the public domain relating to a specific topic and
structuring it into a framework consistent with thinking from a risk assessment point of
view.
D’Elia et al (1998) present a very useful example of the application of the
geotechnical characterisation system of slope movements for structurally complex clay
soils and stiff jointed clays of the Italian Apennine for the various stages of moveme nt
from pre-failure through to reactivation. From the Italian (research by Picarelli, Esu,
their colleagues and others) and UK perspective (research by Skempton, Chandler,
Potts, their colleagues and others), the bringing together of research on stiff fissured or
jointed clays, clay shales and indurated fine grained soils, such as by Picarelli (2000),
provides for a broader understanding of slope movements and the mechanisms of slope
movement in these clay soil types. What is evident for both regions is the wealth of
research invested over many 10’s of years in areas related to landsliding that has been
required in developing the state of knowledge to its present level of understanding.
Within this framework it is recognised that there may be regionally, or even locally or
geologically, specific influences. Overall though, the developed concepts and research
findings are useful for the broader application in other areas where the research is not as
well developed, notwithstanding the need for recognition of regional differences.
Chapter 2: Literature Review Page 2.5

Another example is the Hong Kong experience of landslides in cuts in deeply


weathered soils, loose fills and natural slopes triggered by prolonged or intense rainfall.
The establishment of the Geotechnical Engineering Office by the Hong Kong
Government in 1977 to address the regional issues of landsliding following the
catastrophic consequences of several relatively small volume slides, has seen a wealth
of information gathered and knowledge developed relating to landslides over 10’s of
years (Brand et al 1984; Brand 1985b; Wong and Ho 1996; for example) for use in
quantitative risk assessment (Wong et al (1997) for example). Some of the outcomes of
this research include priority ranking of slopes for remediation, focus on landslide types
of high risk and high consequence (e.g. flow liquefaction of loose fills (HKIE 1998; Sun
1999)) and development of rainfall related landslide warning systems.
The post failure deformation behaviour of “rapid” landslides such as flow slides and
debris flows in a range of material types from soil through to rock has been an active
area of research for the last 20 to 30 years. Research by eminent researchers such as
Cruden, Sassa, Hungr, Iverson and Hutchinson (Sassa 1988; Hutchinson 1986, 1988;
Hungr 1995; Smith and Hungr 1992; Iverson et al 1997; Corominas 1996a; van Gassen
and Cruden 1989) amongst others has contributed to the state of knowledge on post
failure deformation of these slide types.
The concept of creep is a useful macroscopic model for evaluation of the pre-failure
deformation behaviour of first time slides. The term creep is somewhat misleading in
the context as used above. The mechanical meaning of creep, as described by Picarelli
(2000), “implies viscous movements of a soil mass subjected to a constant effective
stress field characterised by local values of stress smaller than soil strength”. But, as
used above it implies more than this and encapsulates the overall pre-failure
deformation leading up to failure including creep itself (in the true sense of the word)
and deformations due to the progressive development of the shear surface for slopes
where the effective stress field does not change.
Mitchell (1964) developed the concept of creep from rate process theory. The
various phases of creep under constant deviatoric stress conditions (Figure 2.3) as
described by Singh and Mitchell (1968) are:
• Primary creep – deformation at a decreasing rate of strain with time. This is
essentially the mechanical meaning defined by Picarelli (2000);
• Tertiary creep – deformation at an accelerating rate of strain with time (under
constant effective stress conditions) culminating in creep rupture or failure. This
Chapter 2: Literature Review Page 2.6

incorporates deformations related to the formation of a shear surface that are outside
of the mechanical meaning of creep; and
• Secondary creep – deformation at a constant rate with time. Varnes (1982)
considered secondary creep as the concurrent processes of primary and tertiary
creep.

An example of the use of the creep model to describe the stages of pre-failure
deformation behaviour is shown in Figure 2.4 for the failure in the retained cut slope in
London Clay at Kensal Green (Skempton 1964). The creep concept is also useful for
analysis of the post construction deformation behaviour of embankments.
Another application of the strain rate concepts is the prediction of the time to failure
for first time slides in soil and rock slopes based on the minimum strain rate in the pre-
failure period (Saito 1965). The method is applicable where the overall effective stress
conditions within the slope are near constant.

Primary Tertiary
creep creep
log e
X Rupture

Minimum
e = strain rate
strain rate
Minimum

time or strain
Figure 2.3: Concept of the creep model under constant deviatoric stress conditions.
Chapter 2: Literature Review Page 2.7

Figure 2.4: Monitored deformation behaviour at retained cut slope failure in London
clay at Kensal Green (Skempton 1964)

2.2 DEFORMATION BEHAVIOUR OF EMBANKMENT DAMS


The embankment types considered in this thesis consist of earthfill, rockfill (concrete
face rockfill dams), and zoned earth and earth-rockfill dams. The earthfills, apart from
the core of puddle dams, are rolled and well compacted in most cases, particularly the
zone acting as the main water barrier element. The placement of the rockfill zones
covers a broad range from well compacted to dumped in high lifts.
The general features of the stress-strain relationship of rolled earthfills are well
described in textbooks on soil mechanics and are applicable to describe the deformation
behaviour of rolled earthfills. Factors such as the degree of over-consolidation,
permeability, soil structure, matric suction and the rate, magnitude and direction of
loading affect the strength and compressibility properties of an earthfill. Some of these
factors are affected by:
• The soil type, including its mineralogy, gradation and plasticity;
• The degree of compaction; and
• The moisture content at placement relative to Standard optimum moisture content.
Chapter 2: Literature Review Page 2.8

For rockfills, field observations (Mori and Pinto 1988; and others) and the results of
large scale laboratory tests (Marsal 1973; Marachi et al 1969; and others) indicate the
compressibility properties of rockfill are affected by:
• Degree of compaction of the rockfill;
• Applied stress conditions and stress path;
• Particle shape and particle size distribution; and
• Intact strength of the rock used as rockfill.

An important aspect in the evaluation of the deformation behaviour of rockfill in


embankment dams is its susceptibility to collapse compression on wetting (Nobari and
Duncan 1972a; Marsal 1973; Justo 1991; Alonso and Oldecop 2000). Collapse
compression of earthfills is equally important, but in rolled earthfills its occurrence is
much less frequent.
For embankment dams, the components of deformation are those that occur as a
result of changes in effective stress conditions (during construction, on impoundment
and due to reservoir fluctuation), changes in total stress (for wet placed earthfill cores in
zoned embankments), on saturation or wetting (e.g. collapse compression) and the on-
going time dependent or creep type deformations. Predictive methods are typically
divided into the three components of deformation during construction, on first filling
and long-term post construction (or post first filling).
Most predictive methods cover the deformation behaviour of one or two of these
components. Finite element analyses have been used (Saboya and Byrne 1993;
Kovacevic 1994; Naylor et al 1997; Dounias et al 1996, amongst others) for analysis of
deformation due to changes in stress conditions, mostly for deformation during
construction and first filling, but also for deformation under large drawdown. Empirical
predictive methods, which are usually based on historical performance of embankments,
are more generally available for assessment of post construction deformation (Sowers et
al 1965; Clements 1984; Dascal 1987; Charles 1986; Soydemir and Kjœrnsli 1975;
amongst others) some of which incorporate the deformation on first filling, although
methods are available to assist with the assessment / prediction of deformation during
construction (Poulos et al 1972; Penman et al 1971; Gould 1953; Pinto and Marques
Filho 1998; Giudici et al 2000).
Chapter 2: Literature Review Page 2.9

2.3 WHERE THIS R ESEARCH F ITS IN


The main focus of this thesis is toward improving the understanding and prediction of
the deformation behaviour of landslides and embankment dams for the classes of slope
and material types considered. Within the context of the geotechnical characterisation
system of slope movements, this thesis concentrates on the deformation related aspects
within the framework rather than providing a comprehensive consideration of all
components for certain slope types. The outcomes are in the form of guidelines and/or
methods for prediction or for evaluation of the deformation behaviour: pre and post
failure deformation behaviour for landslides, and more generally the deformation
behaviour of embankment dams. They are geared toward practical usage by consulting
geotechnical engineers, geologists and engineering geologists within an overall risk
assessment framework.
For embankment dams the database represents one of the largest put together for the
types of embankments considered. The outcomes are geared toward providing dam
owners and their consultants with a framework (developed from analysis of the
database) within which they are able to evaluate and/or predict the deformation
behaviour of their structure.
Chapter 3: “Rapid” landslides from failures in soil slopes Page i

TABLE OF CONTENTS

3.0 “RAPID” LANDSLIDES FROM FAILURES IN SOIL


SLOPES ....................................................................................... 3.1
3.1 OUTLINE OF THIS C HAPTER ............................................................................. 3.1
3.2 DATABASE OF CASE S TUDIES ANALYSED ........................................................ 3.2
3.2.1 Flow Slides in Coal Mine Waste Spoil Piles and Stockpiled Coal............ 3.5
3.2.2 Landslides from Failures in Cut, Fill and Natural Slopes, Hong
Kong .......................................................................................................... 3.7
3.2.3 Flow Slides from Tailings Dam Failures .................................................. 3.9
3.2.4 Flow Slides from Failures in Hydraulic Fill Embankment Dams............. 3.9
3.2.5 Flow Slides from Failures in Sub-Aqueous Constructed Fills.................. 3.9
3.2.6 Flow Slides from Failures in Submarine Slopes ..................................... 3.12
3.2.7 Flow Slides in Sensitive Clays................................................................. 3.12
3.3 THE MECHANICS OF SHEARING OF CONTRACTIVE GRANULAR SOILS ............ 3.15
3.3.1 Definitions ............................................................................................... 3.15
3.3.2 Characteristics of Contractive and Dilative Soils Sheared Under
Monotonic Load Conditions.................................................................... 3.16
3.3.3 Field Methods for Evaluation of Flow Liquefaction Potential ............... 3.22
3.3.4 Effects of Non-plastic Fines and Gravels based on Flow Liquefaction
Potential .................................................................................................. 3.30
3.3.5 Residual Undrained Shear Strength of Flow Liquefied Soils.................. 3.32
3.4 MECHANICS OF FAILURE OF FLOW SLIDES .................................................... 3.39
3.4.1 Mechanics of Development of Flow Sliding............................................ 3.39
3.4.2 Flow Slides in Granular Stockpiles......................................................... 3.43
3.4.3 Flow Slides from Failures in Sensitive Clays.......................................... 3.49
3.4.4 Flow Slides in Loose Silty Sand Fills, Hong Kong..................................3.54
3.4.5 Tailings Dams..........................................................................................3.57
3.4.6 Hydraulic Fill Embankment Dams.......................................................... 3.60
3.4.7 Submarine Slopes and Slopes in Sub-Aqueous Fills ............................... 3.62
3.5 MECHANICS OF FAILURE IN DILATIVE SOILS LEADING TO “RAPID”
SLIDING .........................................................................................................3.64
3.5.1 Slope Types of Failures in Dilative Soils Leading to “rapid” Sliding.... 3.64
3.5.2 Slides Through Soil Mass – Dilative Failures......................................... 3.65
3.5.3 Defect Controlled Landslides.................................................................. 3.79
Chapter 3: “Rapid” landslides from failures in soil slopes Page ii

3.6 SUMMARY OF THE C HARACTERISTICS AND MATERIAL TYPES OF “RAPID”


LANDSLIDES ..................................................................................................3.83
3.6.1 Flow Slides in Contractant Soils Susceptible to Flow Liquefaction ....... 3.83
3.6.2 Characteristics and Identification of Conditions Under Which
Failures in Dilative Soils Result in “Rapid” Landsliding ...................... 3.88
3.7 PRE-FAILURE DEFORMATION BEHAVIOUR ...................................................... 3.89
3.7.1 Flow Slides ..............................................................................................3.90
3.7.2 “Rapid” Landslides from Slope Failures in Dilative Soils ..................... 3.97
3.7.3 Summary of Pre Failure Deformation Behaviour................................... 3.99
3.8 POST-FAILURE DEFORMATION BEHAVIOUR – R EVIEW OF THE
LITERATURE ................................................................................................3.102
3.8.1 Factors Affecting Travel Distance and Velocity ...................................3.102
3.8.2 Velocity of the Slide Mass .....................................................................3.104
3.8.3 Empirical Methods for Assessment of Travel Distance.........................3.107
3.9 POST-FAILURE DEFORMATION – DEVELOPED METHODS FOR ASSESSMENT
OF TRAVEL DISTANCE FROM THE D ATABASE ..............................................3.112
3.9.1 Travel Distance Versus Slide Volume and Initial Slide Classification .3.113
3.9.2 Travel Distance Versus Slide Volume, Degree of Confinement of the
Travel Path and Initial Slide Classification..........................................3.116
3.9.3 Travel Distance Versus Volume, Down-slope Angle and Slide Type
from Failures in Dilative Soils ..............................................................3.124
3.9.4 Travel Distance Versus Volume, Down-slope Angle and Slide Type
from Failures in Contractile Soils.........................................................3.137
3.9.5 Summary of Methods for Predicting Travel Distances .........................3.150
3.10 POST-FAILURE NUMERICAL MODELLING - METHODS AND RESULTS ...........3.152
3.10.1 “Rapid” Landsliding in Hong Kong .....................................................3.155
3.10.2 “Rapid” Landslides from Coal Mine Waste Spoil Pile Failures in
British Columbia ...................................................................................3.158
3.10.3 Flow Slides from Tailings Dams Failures.............................................3.160
3.11 CONCLUSIONS .............................................................................................3.162
Chapter 3: “Rapid” landslides from failures in soil slopes Page iii

LIST OF TABLES

Table 3.1: Summary of database on “rapid” landslides. ................................................ 3.3


Table 3.2: Summary of data sources for “rapid” landslides from coal mine waste
spoil piles and coal stockpiles................................................................................. 3.5
Table 3.3: Summary of landslide cases analysed from Hong Kong .............................. 3.8
Table 3.4: Summary of case studies of “rapid” landslides from tailings dams ............ 3.10
Table 3.5: Summary of failure case studies in hydraulic fill embankment dams ........ 3.11
Table 3.6: Summary of failure case studies in sub-aqueous constructed fill slopes .... 3.11
Table 3.7: Summary of “rapid” flow slides in natural submarine slopes..................... 3.13
Table 3.8: Summary of flow slides in sensitive clays .................................................. 3.14
Table 3.9: SPT blow count ((N1)60) yield strength and residual strength correction
factors for fines content (Seed et al 1985; Seed 1987)..........................................3.35
Table 3.10: Features of flow slides in loose fill from Hong Kong...............................3.56
Table 3.11: Summary of properties of liquefaction susceptible materials from flow
slides in hydraulic fill embankments..................................................................... 3.61
Table 3.12: Summary of material types of dilative slides of debris in natural
slopes that developed in “rapid” debris slides and debris flows. .......................... 3.75
Table 3.13: Source area slope angle (α 1 ) of dilative slides in natural slopes that
developed into “rapid” debris flows and debris slides.......................................... 3.77
Table 3.14: Failures through the soil mass in dilative soils that developed into
“rapid” landslides, from Hong Kong..................................................................... 3.78
Table 3.15: Features of defect controlled compound slides that developed into
“rapid” landslides from Hong Kong...................................................................... 3.82
Table 3.16: Features of defect controlled translational slides that developed into
“rapid” landslides from Hong Kong...................................................................... 3.82
Table 3.17: Summary of pre-failure observations in coarse-grained loose fills .......... 3.92
Table 3.18: Measured slope deformations adjacent to failures in the Ottawa area
(Mitchell and Eden 1972)...................................................................................... 3.95
Table 3.19: Summary of characteristics of pre-failure deformation of slope
failures that developed into “rapid” landslides ...................................................3.100
Table 3.20: Velocity of the slide mass for “rapid” landslides from failures in soil
slopes (mostly from case studies analysed).........................................................3.105
Table 3.21: Velocity of slide mass for debris slides and debris flows.......................3.106
Chapter 3: “Rapid” landslides from failures in soil slopes Page iv

Table 3.22: Regression analysis of H/L (Equation 3.11) versus landslide volume
(Corominas 1996a)..............................................................................................3.111
Table 3.23: Summary of empirical correlation coefficients (Equation 3.11) for
debris flows by Corominas (1996a) ....................................................................3.114
Table 3.24: Empirical correlation coefficients for power law fit of travel distance
angle to slide volume (Equation 3.11, H L = AV B ) for unconfined “rapid”
landslides.............................................................................................................3.122
Table 3.25: Empirical correlation coefficients for power law fit of travel distance
angle to slide volume (Equation 3.11, H L = AV B ) for confined and partly
confined “rapid” landslides.................................................................................3.124
Table 3.26: Summary of empirical correlations for flow slides in coal waste spoil
piles in British Columbia ....................................................................................3.139
Table 3.27: Statistical summary of H/L to tan a 2 correlation for flow slides in
coarse-grained coal mine waste spoil piles.........................................................3.143
Table 3.28: Summary of slide properties of flow slides in hydraulic fill
embankments.......................................................................................................3.149
Table 3.29: Mean and standard deviation of H/L for several types of slopes giving
“rapid” landslides................................................................................................3.153
Table 3.30: Summary of recommended methods for prediction of H/L (tangent of
the travel distance angle).....................................................................................3.154
Table 3.31: Summary of “rapid” landslides from Hong Kong analysed by Hungr
Geotechnical Research (1998) and Ayotte and Hungr (1998)............................3.156
Table 3.32: Summary of results of numerical modelling using DAN for “rapid”
landslides in Hong Kong (Hungr Geotechnical Research 1998; Ayotte and
Hungr 1998) ........................................................................................................3.157
Chapter 3: “Rapid” landslides from failures in soil slopes Page v

LIST OF FIGURES

Figure 3.1: Typical coal waste spoil pile profile in British Columbia (Hungr et al
1998) ....................................................................................................................... 3.6
Figure 3.2: Steady state line of Toyoura sand (Ishihara 1993) .................................... 3.16
Figure 3.3: Undrained behaviour of states on the contractive side of the steady
state line (Ishihara 1993)....................................................................................... 3.17
Figure 3.4: Undrained behaviour from states close to steady state and on dilative
side (a) stress-strain plot and (b) effective stress paths (Ishihara 1993) ............... 3.17
Figure 3.5: Schematic diagram showing different types of behaviour for undrained
stress paths of samples at the same void ratio and different initial confining
pressures (Yamamuro and Lade 1998).................................................................. 3.18
Figure 3.6: Definition of the state parameter (Been and Jefferies 1985). .................... 3.20
Figure 3.7: Collapse surface concept, defined by the locus of the peak deviator
stress from undrained triaxial samples tested at the same void ratio and
different initial effective mean normal stress (Sladen et al 1985a)....................... 3.20
Figure 3.8: Correlation for cyclic resistance ratio (CRR) based on corrected SPT
blow count, (N1)60, for clean sands under level ground conditions (Robertson
and Wride 1997).................................................................................................... 3.23
Figure 3.9: Corrected SPT blow count, (N1)60, versus vertical effective over-
′ ) for saturated non-gravelly silty sand deposits that have
burden pressure (σ vo
experienced large deformations and were triggered by earthquake (Baziar and
Dobry 1995) .......................................................................................................... 3.24
Figure 3.10: Flow liquefaction and no-flow bounds in terms of (a) field measured
SPT blow count and (b) CPT qc value from flow liquefaction case histories
(static and earthquake triggered) and laboratory undrained testing (Ishihara
1993). .................................................................................................................... 3.25
Figure 3.11: Influence of void ratio range on flow potential of sandy soils for (a)
round-grained sands and (b) angular grains (Cubrinovski and Ishihara 2000)..... 3.26
Figure 3.12: Relationship between relative density, Dro , at the threshold void ratio,
eo, and void ratio range (e max – e min) for sandy soils (Cubrinovski and Ishihara
2000). .................................................................................................................... 3.27
Figure 3.13: Relationship between the slope of the steady state line, λ , and void
ratio range for (a) round-grained sands and (b) angular grained sands
(Cubrinovski and Ishihara 2000)........................................................................... 3.28
Chapter 3: “Rapid” landslides from failures in soil slopes Page vi

2
Figure 3.14: Correlation between N1 Dr (N1 = SPT blow count corrected for
over-burden pressure, Dr = relative density) and void ratio range for
cohesionless soils (Cubrinovski and Ishihara 2000). ............................................ 3.28
Figure 3.15: Field measured SPT blow count boundaries differentiating flow
liquefaction no-flow conditions for void ratio range of 0.3 to 0.6 for (a)
round-grained sands and (b) angular sands (Cubrinovski and Ishihara 2000)...... 3.29
Figure 3.16: Flow characterisation of sand based on field measured SPT blow
count showing liquefaction flow no-flow boundary, and region of flow at zero
residual undrained strength (Cubrinovski and Ishihara 2000) .............................. 3.30
Figure 3.17: Increase in liquefaction potential (as measured by relative density)
with increasing content of non-plastic fines (Lade and Yamamuro 1997) ........... 3.31
Figure 3.18: Fording sandy gravel sample (a) undrained triaxial compression test
results and (b) consolidation and steady states (Dawson et al 1998). ................... 3.32
Figure 3.19: Relationship between equivalent clean sand SPT blow count, (N1)60-
CS , and residual undrained strength from field case studies (Seed and Harder
1990) ..................................................................................................................... 3.34
Figure 3.20: Relationship between undrained critical strength ratio and equivalent
clean sand SPT blow count, (N1)60-CS (Stark and Mesri 1992).............................. 3.35
Figure 3.21: Residual undrained shear strength, Sr, to vertical effective over-
burden pressure for silty sand deposits (≥ 10% fines), triggered by
earthquake, that have experienced large deformation (Baziar and Dobry 1995).. 3.36
Figure 3.22: Flow chart for assessment of residual undrained shear strength of
liquefied soil (NSF 1998)...................................................................................... 3.38
Figure 3.23: Idealised mechanism for initiation of flow slides.................................... 3.40
Figure 3.24: Definition of undrained brittleness index, IB, in strain weakening soil. .. 3.41
Figure 3.25: Effective stress states on some planes (submarine slopes) located
within the region of potentially instability (Lade 1992). ...................................... 3.42
Figure 3.26: Collapse model for mine waste dumps (Dawson et al 1998) .................. 3.43
Figure 3.27: Particle size distribution of coal mine waste and coking coal
susceptible to static liquefaction ........................................................................... 3.44
Figure 3.28: Particle size distribution of sensitive clays .............................................. 3.50
Figure 3.29: Frequency distribution of landslides in Quebec (Tavenas 1984) ............ 3.52
Figure 3.30: Slope height versus slope inclination for unstable slopes in eastern
Canada (Tavenas 1984)......................................................................................... 3.52
Figure 3.31: Retrogressive landslide in sensitive clay (Leroueil et al 1996) ............... 3.54
Figure 3.32: Peak shear strength from undrained triaxial compression tests of
contractive fill samples (HKIE 1998) ................................................................... 3.57
Chapter 3: “Rapid” landslides from failures in soil slopes Page vii

Figure 3.33: Effect of partial saturation on steady state line in void ratio – mean
effective stress space (HKIE 1998)....................................................................... 3.58
Figure 3.34: Particle size distribution of tailings from flow slide case studies............ 3.59
Figure 3.35: Particle size distributions of hydraulic fills from flow slide case
studies.................................................................................................................... 3.61
Figure 3.36: Particle size distributions of soil types from submarine slopes and
sub-aqueous fills susceptible to liquefaction and flow sliding.............................. 3.62
Figure 3.37: Hypothetical model for transformation of dilative materials from
rigid to fluid behaviour, path Di to Df (Fleming et al 1989).................................. 3.68
Figure 3.38: Relationship between peak hourly rainfall, landsliding and severity of
consequence for Hong Kong (Brand et al 1984)................................................... 3.70
Figure 3.39: Relationship between severity of landsliding and rainfall intensity for
Hong Kong (Brand 1985b).................................................................................... 3.70
Figure 3.40: Source area particle size distributions of dilative slides of debris in
natural slopes......................................................................................................... 3.73
Figure 3.41: Distribution of head slope angle (source area slope angle) for slides
of debris in natural terrain that developed into “rapid” landslides, Hong Kong
(Evans et al 1997).................................................................................................. 3.79
Figure 3.42: Cross section through the failure and “rapid” landslide at Fei Tsui
Road, Hong Kong (GEO 1996a)........................................................................... 3.81
Figure 3.43: Particle size distributions of material types susceptible to liquefaction
and flow sliding. .................................................................................................... 3.84
Figure 3.44: Comparison of flow liquefaction no-flow boundaries (in terms of
SPT (N1)60) for sands and silty sands from monotonic laboratory undrained
tests and earthquake triggered field cases (after Cubrinovski and Ishihara
2000). ....................................................................................................................3.87
Figure 3.45: Pre-failure deformation of two separate failures at the Clode waste
dump, Fording Coal, British Columbia (Campbell and Shaw 1978)....................3.93
Figure 3.46: Measured deformations in a natural slope, Ottawa River (Mitchell
and Eden 1972) ..................................................................................................... 3.94
Figure 3.47: Deformation and pore pressure ratio of an induced slope failure
(Mitchell and Williams 1981) ............................................................................... 3.96
Figure 3.48: Sliding block model showing kinetic energy lines for different
rheological models (Golder Assoc. 1995)...........................................................3.106
Figure 3.49: Schematic profile of a slope failure showing travel distance and
travel distance angle............................................................................................3.108
Chapter 3: “Rapid” landslides from failures in soil slopes Page viii

Figure 3.50: Travel distance from the toe of the cut, R, versus cut slope height, Hs,
for failures in cut slopes from Lantau Island, Hong Kong (Wong and Ho
1996) ...................................................................................................................3.109
Figure 3.51: Landslide volume versus tangent of the travel distance angle for 204
landslides (Corominas 1996a).............................................................................3.110
Figure 3.52: Ratio H/L versus volume for landslides that break-up on sliding
(Finlay et al 1999) ...............................................................................................3.111
Figure 3.53: Ratio H/L versus volume for all slides from database...........................3.114
Figure 3.54: H/L ratio versus volume relationships for Corominas (1996a) data on
unconfined debris flows. .....................................................................................3.115
Figure 3.55: H/L ratio versus volume relationships for Corominas (1996a) data on
confined debris flows. .........................................................................................3.115
Figure 3.56: Travel distance beyond source area versus volume for unconfined
travel paths. .........................................................................................................3.117
Figure 3.57: Travel distance beyond source area versus volume for confined and
partly confined travel paths. ................................................................................3.117
Figure 3.58: Comparison of trendlines of H/L versus slide volume differentiated
on travel path confinement for “rapid” landslides from failures in dilative
soils, from Hong Kong. .......................................................................................3.118
Figure 3.59: H/L versus volume for “rapid” landslides with unconfined travel
paths. ...................................................................................................................3.120
Figure 3.60: H/L ratio versus volume for “rapid” landslides from failures in
dilative soils with unconfined travel paths..........................................................3.120
Figure 3.61: H/L ratio versus volume for “rapid” landslides from failures in
contractive soils (flow slides) with unconfined travel paths. ..............................3.121
Figure 3.62: H/L versus volume for “rapid” landslides with confined travel paths...3.123
Figure 3.63: H/L versus volume for “rapid” landslides with partly confined travel
paths. ...................................................................................................................3.123
Figure 3.64: H/L versus slide volume for cut slope cases studies from Hong Kong .3.126
Figure 3.65: H/L normalised against tangent of the cut slope angle versus volume
for cut slopes in Hong Kong failing onto a near horizontal slope at the toe.......3.127
Figure 3.66: Idealised effect of volume on travel distance for cut slope failures of
(a) small volume compared with (b) significantly larger volumes. ....................3.127
Figure 3.67: Classification of slope categories (a) Type 1 – failure at top of cut
slope, and (b) Type 2 – failure encompassing the full cut height. ......................3.128
Figure 3.68: Definition of down-slope angle below source area, a 2..........................3.132
Chapter 3: “Rapid” landslides from failures in soil slopes Page ix

Figure 3.69: H/L versus tangent of the down-slope angle, a 2, for unconfined travel
paths, failures in natural slopes from Hong Kong...............................................3.133
Figure 3.70: H/L versus tangent of the down-slope angle, a 2, for partly confined
travel paths, failures in natural slopes from Hong Kong.....................................3.134
Figure 3.71: H/L versus tangent of the down-slope angle, a 2, for confined travel
paths, failures in natural slopes from Hong Kong...............................................3.134
Figure 3.72: Natural slopes (Hong Kong). Trendlines for H/L versus tangent of
the down-slope angle, a 2, for all types of travel path. .........................................3.135
Figure 3.73: H/L versus slide volume for flow slides in loose silty sand to sandy
silt fills.................................................................................................................3.137
Figure 3.74: H/L versus slide volume for flow slides in coarse-grained coal mine
waste spoil piles. .................................................................................................3.142
Figure 3.75: H/L versus the tangent of the down-slope angle below the toe of the
spoil pile, a 2; flow slides in coarse-grained coal mine waste spoil piles. ...........3.142
Figure 3.76: H/L versus volume for retrogressive “rapid” slides...............................3.146
Figure 3.77: Distance of retrogression versus liquidity index for slides in sensitive
clays.....................................................................................................................3.148
Figure 3.78: Distance of retrogression based on stability number for landslides in
sensitive clays (Mitchell and Markell 1974).......................................................3.148
Figure 3.79: Ratio of retrogression distance to slide depth versus stability number .3.149
Figure 3.80: DAN results of frictional back-analysis models of “rapid” landslides
from Hong Kong (data from Hungr GR (1998) and Ayotte and Hungr (1998)) 3.158
Figure 3.81: Back calculated bulk friction angle from DAN analysis of flow slides
from coal mine waste spoil pile failures in British Columbia (Golder Assoc.
1995) ...................................................................................................................3.159
Figure 3.82: Simple sliding block model for analysis of tailings strength (Blight
1997) ...................................................................................................................3.160
Figure 3.83: Variation of shear strength with moisture content for gold tailings
(Blight 1997) .......................................................................................................3.161
Figure 3.84: Model for flow of liquefied tailings (Jeyapalan et al 1983a).................3.162
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.1

3.0 “RAPID” LANDSLIDES FROM FAILURES IN SOIL


SLOPES

3.1 OUTLINE OF THIS CHAPTER


“Rapid” landslides from failures in soil slopes have mass destructive capabilities;
resulting in loss of life, destruction of property and damage to the natural environment.
Examples are Aberfan, South Wales in 1966 (failure of coal waste stockpile resulting in
144 deaths); Stava, Italy in 1985 (failure of tailings dam resulting in some 268 deaths);
earthquake triggered failures in loessic soils in Gansu Province, China in 1920 (killed
some 100,000 people); loss of life from relatively small volume landslides in Hong
Kong; and the relatively small volume landslide at Thredbo Village, Australia in 1997
resulting in 18 deaths.
Most of this chapter is dedicated to presenting the results from analysis of a database
of some 350 “rapid” landslides. The case studies, in most cases, are of landslides that
post-failure have reached travel velocities within the rapid, very rapid and extremely
rapid IUGS (1995) velocity categories (Table 1.1). This is as opposed to debris flows
that have evolved through channel erosion, which are not considered here. In most
cases the initial landslide has been “statically” triggered as opposed to “dynamically”
triggered (e.g. by blasting or earthquake), but several dynamically triggered landslides
are included in the database. The chapter also draws on the published literature on
failures in soil slopes that developed into “rapid” landslides, including areas of the
literature on cyclic liquefaction of granular soils.
The main objectives of this chapter are centred on:
• The characteristics of soils and slope conditions within which “rapid” landslides
develop, including the identification of soils susceptible to liquefaction and potential
flow behaviour;
• Identification of the triggers and failure mechanism/s of slides in soil slopes that
develop into “rapid” landslides post failure;
• Review and refinement of empirical methods for prediction of travel distance;
• Review of numerical models to predict travel of the landslide, what models are
available and under what conditions are they applicable.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.2

The database of “rapid” landslides from failures in soil slopes assembled for this
study is briefly summarised (Section 3.2). This is followed by a review of the literature
on the mechanics of shearing of contractive granular soils (Section 3.3), which leads
into a discussion on the mechanics of failure for the slope and material classes
considered; Section 3.4 for flow slides in soils that are potentially contractile on
shearing and Section 3.5 for landslides in soils that are dilative on shearing (Section 3.6
is a summary). The analysis of the pre-failure (Section 3.7) and post-failure
deformation behaviour (Sections 3.8 to 3.10) is then presented, which includes methods
developed for travel distance estimation for several classes of slope types considered.

3.2 DATABASE OF CASE S TUDIES ANALYSED


The database comprises some 350 “rapid” landslides from failures in predominantly soil
slopes from published sources. Table 3.1 presents a summary of material types of the
landslides sorted in terms of the initial landslide type, either initially in contractive or
dilative soil types. For some 100 individual cases, reasonably detailed data was
available from either reports or published conference and journal papers. The
remainder of the data was gathered from other sources including Corominas (1996a),
Golder Assoc. (1995), Siddle et al (1996), Hutchinson (1988), Wong and Ho (1996),
Sun (1999) and Edgars and Karlsrud (1983). Tabulated information on almost 200 of
the “rapid” landslides from failures in soil slopes is presented in Section 1 of Appendix
B.
The travel classification of most of the “rapid” landslides is debris flow or flow
slide. There are only several that would be classified as debris slides because most
slides that initiate in dilative soils tend to break-up and flow rather than move as a
relatively intact slide mass on a defined basal sliding plane.
The database of debris flows from Corominas (1996a) incorporates a number of
slides initially within rock that developed into debris flows. The sub-groups chalk cliffs
and sturztroms are also slides initially within rock that developed into flow slides and
debris flows. The purpose of including these events is for comparison with debris flows
initiated in soil slopes.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.3

Table 3.1: Summary of database on “rapid” landslides.


Initial Landslide Material Sub-group No. Volume Range Material Description References
Type Type (m3)
Road and building 16 50 to 10,500 Loose dumped silty sands to sandy silts Wong and Ho (1996), Sun
fills, Hong Kong with low clay content. (1999), GEO reports
Coal waste spoil piles 56 3000 to 8x106 Loose dumped sandy gravels (British Refer Appendix B, Section 1.
Columbia and South Wales)
Coking coal stockpiles 9 850 to 16,000 Sandy gravels to gravelly sands, low Refer Appendix B, Section 1.
specific gravity (Hay Point)
Loose Fills Hydraulic fill 5 16,000 to 8x106 Hydraulically placed silty and sandy Refer Appendix B, Section 1.
Slides in Contractile embankment dams*2 soils, stratified (1 case is of loose
Soils dumped sand).
6
Tailings dams 9 20,000 to 3x10 Hydraulically placed sand to clay sized Refer Appendix B, Section 1.
soils, stratified.
Sub-aqueous hydraulic 4 4,000 to 150,000 Hydraulically placed and dumped sands Refer Appendix B, Section 1.
and dumped fills and silty sands.
6
Sensitive clays 12 100,000 to 55x10 Leached marine deposits, highly Refer Appendix B, Section 1.
Natural Soils
sensitive, clayey silts to silty clays.
Slopes 12
Submarine slopes 22 6,000 to 1x10 Mainly in sands and silts. Edgars and Karlsrud (1983).
*1 Inclusive of defect controlled slides, slides of debris and slides through the soil mass
*2 Includes 4 flow slides in hydraulic fill embankment dams and 1 flow slide in a loose dumped sand fill embankment dam
GEO = Geotechnical Engineering Office, Hong Kong Government
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.4

Table 3.1 (cont.): Summary of database on “rapid” landslides (sheet 2 of 2)


Initial Landslide Material Sub-group No. Volume Range Material Description References
Type Type (m3)
Road and building 61 4 to 52,000 Colluvium, residual soils and saprolite. Wong and Ho (1996), Ayotte
Cut Slopes cuts, Hong Kong Silty sands to sandy silts with gravel to and Hungr (1998), GEO
boulders, low clay content. reports
Road and building 10 40 to 500 Mainly silty sands to sandy silts, low Sun (1999), GEO reports
Fill Slopes fills, Hong Kong clay content.
(dilative) Washouts of road fills, 19 40 to 4000 Silty sands to sandy silts with gravel to Wong and Ho (1996), Sun
Hong Kong boulders, low clay content. (1999)
Natural slopes in 27 20 to 26,000 Mostly in colluvium, some in residual Refer Appendix B, Section 1.
Slides in Dilative granitic and volcanic soils and saprolite. Silty sands to sandy
Soils*1 terrain, Hong Kong silts with gravel to boulders, low clay
Natural Soil
content.
and Rock
Chalk talus 4 1,200 to 14,000 - Hutchinson (1988)
Slopes
Chalk cliffs 13 25,000 to 1.3x106 Rock slope failures in chalk cliffs of Hutchinson (1988), Golder
UK. Assoc. (1995)
Kaolinised granite 1 6000 Failure in kaolinised granite Hutchinson (1988)
10
Debris flows Soil and rock slopes 70 36 to 1.2x10 Debris flows emanating from initial Corominas (1996a)
in Soil and slides in soil and rock.
6 9
Rock Sturzstroms 9 10 to 1.5x10 Essentially from initial slides in rock. Hutchinson (1988)
1
* Inclusive of defect controlled slides, slides of debris and slides through the soil mass
*2 Includes 4 flow slides in hydraulic fill embankment dams and 1 flow slide in a loose dumped sand fill embankment dam
GEO = Geotechnical Engineering Office, Hong Kong Government
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.5

3.2.1 FLOW SLIDES IN COAL MINE WASTE SPOIL PILES AND STOCKPILED COAL

“Rapid” flow slides from failures in coal mine waste spoil piles and coal stockpiles have
been reported by a number of published sources. Data for some 67 flow slides has been
analysed from three geographic regions as summarised in Table 3.2, all of associated
with the coal industry. Flow slides in granular stockpiles have been reported in other
geographic regions and from other than the coal industry (Bishop 1973), however these
three regions represent the better sources of available published information particularly
with respect to general information on material properties, topography and climate.
For most of the case studies within the database only limited data was obtained from
the literature. Detailed information was available for several case studies including the
Aberfan tip failure (Tip 7) in 1966, several other flow slides from South Wales and
several flow slides from British Columbia. Information on the individual slides is given
in Section 1 of Appendix B.

Table 3.2: Summary of data sources for “rapid” landslides from coal mine waste spoil
piles and coal stockpiles
Geographic Material No. of Data Sources
Region Description *1 Slides
British Coal waste - loose end 40 Data sourced mainly from Golder Assoc.
Columbia, dumped spoil piles of (1995). Detailed information on 3 failures
Canada sandy gravels to boulder (Dawson 1994; Dawson et al 1998).
size
South Wales, Coal waste - dumped, 18 Limited data on most case studies from
UK mainly sandy gravels with Siddle et al (1996), data for some slides
trace to some silt sized from Knox (1927) and Bishop (1973).
fines Detailed information and data on Aberfan
and several other failures from HMSO
(1967) and associated technical reports.
Hay Point, Loose stockpiled coking 9 Limited data for field case studies from
Australia coal, low specific gravity, Eckersley (1985, 1986, 1990). Detailed
typically gravelly sands to laboratory data and laboratory scale tests
sandy gravels with trace to from Eckersley (1986, 1990).
some silt fines.
Note: *1 “trace” is less than 5% by weight, and “some” is more than 5% and less than 12% by weight.

Dumping operations and material properties vary between the three geographic regions.
In British Columbia the coal waste is placed by end dumping from haul trucks at the
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.6

edge of the tipping face creating a flat platform at the crest and steep tipping face with
average slopes of 36 to 38 degrees (Figure 3.1). The waste is tipped onto relatively
steep foundation slopes, typically from 10 up to 35 degrees. The waste itself varies
widely from cobble to boulder size strong sandstone rock fragments to finer fractions
(sandy gravels) derived from weaker mudstones and re-handle (materials handled or
moved more than once) with silt contents of less than 5 percent. The stockpile profile
(Figure 3.1) generally comprises layers of the finer waste parallel to the tipping face
within the coarser sandstone waste and a basal zone of coarse waste formed by
segregation during tipping.

Figure 3.1: Typical coal waste spoil pile profile in British Columbia (Hungr et al 1998)

Most of the “rapid” landslides from failures in the coal mine waste spoil piles of South
Wales, as described by Siddle et al (1996), were located within the central to eastern
portion of the coalfield where the hills are capped by the massive Pennant Sandstone
formation and the topography is characterised by greater relief and steeper hill-slopes
than the western part of the coalfield. The method of disposal of the spoil involved
deposition onto the active steep face of the stockpile using methods including tipping
from trams, aerial ropeways and various tippler mechanisms. The waste itself is
predominantly derived from shale with some coal and finer coal washery waste (post
1930’s).
Hay Point, Australia is a shipping port for coking coal. Stockpiles of the coking
coal are typical formed by rail-mounted stackers to a maximum height of about 14 m
onto a flat platform area (Eckersley 1985). Bulldozers are sometimes used to extend the
stockpiles beyond the reach of the stackers. The coking coal itself has a low specific
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.7

gravity (approximately 1.35 t/m3) and classifies as a sandy gravel to gravelly sand with
trace to some silt sized fines.

3.2.2 LANDSLIDES FROM FAILURES IN CUT, FILL AND NATURAL SLOPES, HONG KONG

Hong Kong is particularly susceptible to the development of “rapid” landslides from


slope failures. Its steep topography, tropical climate, geology and intense development
are considered to be the dominant factors for the prevalence of “rapid” landslides.
Debris slides and debris flows in natural slopes are numerically the predominant type of
“rapid” landslide, but “rapid” slides developed from failures in loose fill slopes and
steep cut slopes are economically and from a loss of life viewpoint the most important.
Most failures are typically of shallow depth and are triggered during seasonal heavy
rainfall as a result of a reduction in suction due to infiltration and/or development of
transient high ground water pressures.
As part of this study on “rapid” landslides, forty-three landslides from Hong Kong
have been analysed in detail, the main source of information being specific reports on
the slide by the Geotechnical Engineering Office (GEO), its consultants or the Hong
Kong Government (prior to 1977) supplemented with additional information from other
sources (Hungr Geotechnical Research 1998; Morgenstern 2000; Irfan and Woods
(1998); Irfan (1997); Ayotte and Hungr 1998). Table 3.3 summaries the types of slope
failures analysed that developed into “rapid” landslides. Further details on each case
study are given in Section 1 of Appendix B. A slide volume of 20 m3 was established
as a minimum for detailed analysis.
The term “flow slide” has only been applied to fill slope failures where the failure
mechanism was considered most likely due to shear induced flow liquefaction of the
near saturated soil structure. The assessment of the flow liquefaction potential of a fill
was based firstly on measured density results (if available) and then on the type of
placement and age of the fill. In general, failures in dumped fills of decomposed granite
or volcanics with densities less than 80 to 85% of Standard Maximum Dry Density
(SMDD) have been deemed flow slides based on previous studies of these materials
(HKIE 1998; Gov’t of Hong Kong 1977).
Other studies on “rapid” landslides from Hong Kong have been incorporated into
the overall analysis undertaken. Sources of this information are summarised as follows:
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.8

• Wong and Ho (1996) analysed forty-two landslides, predominantly from failures in


cut slopes, on Lantau Island following the severe rainstorm of 5 November 1993.
They looked at the factors affecting travel distance including; down-slope geometry,
slide volume, mechanism of failure and mode of debris movement.
• Finlay et al (1999), from a database of 1100 landslides in Hong Kong between 1984
and 1993, undertook statistical analysis to assess the correlation of various slope
factors to travel distance for cut slopes, fill slopes, retained fills and boulder falls.
• GEO reports on failures in natural slopes (Wong et al 1996; Franks 1996; Evans et
al 1997). Evans et al (1997) used a regression analysis developed from landslides in
the Tung Chang area of Lantau Island to predict the travel distance angle of
landslides from failures in natural slopes from the south Lantau area reported by
Wong et al (1996).
• Ayotte and Hungr (1998) and Hungr Geotechnical Research (1998) numerically
analysed the travel of some 26 “rapid” landslides from Hong Kong using the DAN
computer program. The cases analysed were mostly failures in natural slopes with
several failures in cut and fill slopes. Their findings are discussed in Section 3.10.1.

Table 3.3: Summary of landslide cases analysed from Hong Kong


Slope No. Failure Mechanism Data Sources Additional
Type Cases (from GEO*1, *2) Cases *3
Fill Slopes 10 6 flow slides in loose fill GEO Report Nos. 52, 88, 89, -
1 dilatant slide through the soil LSR 6/99
mass Gov’t HK (1977)
3 retained slopes, one possibly a
flow slide
Cut Slopes 21 8 strongly defect-controlled GEO Report Nos. 52, 88, 90, 1 with limited
slides (translational and 91, 92, LSR 5/99, LSR 7/99, information
compound). LSR 8/99, ADR 1/95, GR 1/95
12 slides in dilatant residual Halcrow (1998a, 1998b), GEO
soils, colluvium and saprolite, (1994, 1996a), Vail (1972),
some with limited defect control. Knill (1996a), Morgenstern
1 retained cut slope (1994).
Natural 12 11 slides of debris, mostly in GEO Report Nos. 88, 89, 90, 14 with limited
Slopes (and colluvium, of which 8 developed 91, SPR 6/96, SPR 10/96, LSR information. All
quasi- into confined debris flows. 7/99, DN 2/97 developed into
natural or 1 defect-controlled compound GEO (1996b), Knill (1996b), debris flows; 3
modified) slide into confined
debris flows.
Notes: *1 GEO = Geotechnical Engineering Office of Hong Kong Government
*2 A full list of references of the slide specific GEO Reports are given in Appendix B
*2 Additional cases from Ayotte and Hungr (1998)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.9

3.2.3 FLOW SLIDES FROM TAILINGS DAM FAILURES

A total of nine case studies of landslides from tailings dam have been collected from
the literature (Table 3.4). Details of the individual case studies are given in Section 1 of
Appendix B.
Information on earthquake triggered flow slides from tailings dams was also
collected from the published literature (refer Appendix B). However, very little
information was found on the post-failure behaviour of these case studies and they have
therefore not been considered in detail.

3.2.4 FLOW SLIDES FROM FAILURES IN HYDRAULIC FILL EMBANKMENT DAMS

Five case studies of “rapid” landslides from failures in embankment dams constructed
of hydraulically placed or loose dumped sands and silty sands have been analysed
(Table 3.5). Details are given in Section 1 of Appendix B. For the hydraulic fill
embankments, failure was associated with liquefaction of either the hydraulic placed fill
or the foundation. It was evident that the outer denser materials and non-saturated
zones remained virtually intact during sliding, and were carried forward on a liquefied
zone of the looser and saturated materials. An important characteristic of these failures
is the stratified nature of the hydraulically placed materials.
In the case of Wachusett Dam the failure occurred in the upstream shoulder of loose
dumped sand fill during first filling and Olson et al (2000) consider it to have involved
static liquefaction and flow of the loose dumped sand fill.

3.2.5 FLOW SLIDES FROM FAILURES IN SUB-AQUEOUS CONSTRUCTED FILLS

Four case studies of failures that developed into “rapid” flow slides in sub-aqueous
constructed fills have been included in the database (Table 3.6) and all are suspected as
being retrogressive. Details for each case study are given in Section 1 of Appendix B.
All failures were in loose to very loose sands with some to trace silt fines (density index
estimated at 15 to 30%).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.10

Table 3.4: Summary of case studies of “rapid” landslides from tailings dams
Name and Mine / Tailings Embankment Operating Material Description Data Sources
Country Type Type of Materials State at Time of Tailings
Construction of Failure
Stava, Italy Flourite Centreline then Silty sands (SM) In operation Interlayered silty clays Blight (1997)
upstream, cycloned (CL) and silty sands (SM) Chandler and Tosatti (1995)
Morgenstern (2000), Berti et al (1988)
Bafokeng, South Platinum Upstream, Silty sands (SM) with In operation Clayey silts (ML) Fourie (2000), Blight (1997)
Africa mechanical slimes layers Jennings (1979), Rudd (1979)
Midgley (1979)
Merriespruitt, Gold Upstream (?) Sandy soils In operation Sandy silts (ML) Fourie (2000), Blight (1997)
South Africa Wagener et al (1998)
Saaiplass, South Gold Upstream (?) - In operation - Blight (1997)
Africa (3 slides)
Arcturus, Gold Upstream (?) - In operation - Blight (1997)
Zimbabwe
Texas, USA Phosphate, Upstream, Sandier beached In operation Sandy silts (ML) Kleiner (1976)
gypsum tailings mechanical tailings Jeyapalan et al (1983a, 1983b)
Buffalo Creek, Coal Coarse coal waste dumped over existing In operation Sandy silts and sandy Wahler and Assoc. (1973)
USA dammed slimes clays (ML/CL) Johnson (1980)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.11

Table 3.5: Summary of failure case studies in hydraulic fill embankment dams
Name Year of Embankment Material Type/s Location and Timing of Data Sources
Failure Construction Failure
Sheffield Dam, 1925 Hydraulic Silty sands to sandy silts Downstream shoulder, during Seed et al (1969), Seed (1987)
USA earthquake
Lower San 1971 Hydraulic Silty sands to sandy silts Upstream shoulder, during Seed et al (1975), Castro et al (1985, 1992)
Fernando, USA earthquake Gu et al (1993), Baziar and Dobry (1995)
Calaveras Dam, 1918 Hydraulic Sandstone, soft, sluiced Upstream shoulder, during Hazen (1918), Hazen and Metcalf (1918)
USA construction Seed (1987)
Fort Peck Dam, 1938 Hydraulic Sands Upstream shoulder, during ENR (1939a), Middlebrooks (1940)
USA construction Casagrande (1965), Seed (1987)
Wachusett Dam, 1907 Loose dumped Sands Upstream shoulder, during first filling Olson et al (2000)
USA

Table 3.6: Summary of failure case studies in sub-aqueous constructed fill slopes
Name Date of Method of Material Description *1 Timing/Cause of Failure Data Sources
Failure Construction
Nerlerk Sea Berm, 1982 Hydraulically placed sand Sand with trace to some silt, loose (DR = Failure during construction due toSladen et al (1985b)
Canada fill 25 to 35%) localised over-steepening Sladen and Hewitt (1989)
(3 slides) Rogers et al (1990)
Willamette River 18/1/67 Bulldozed sand fill over Sand, medium to coarse grained, very Failure induced by excavation at the Cornforth et al (1974)
Terminal, USA hydraulically placed sand loose (DR = 15%) bulldozed fill to loose to toe of the slope. Flow slide in the
fill medium dense (DR = 20 to 60%). very loose bulldozed fill.
1
Note: * DR = relative density or density index
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.12

3.2.6 FLOW SLIDES FROM FAILURES IN SUBMARINE SLOPES

“Rapid” flow slides from failures in natural submarine slopes were included in the
database for comparative purposes with other flow slides. The twenty-two case studies
used in the post-failure analysis were from Edgars and Karlsrud (1983). Table 3.7 is a
summary of the case studies and further details are given in Section 1 of Appendix B.
The landslides were predominantly of large volume (up to 8 x 1011 m3) and mostly
ancient slides in mainly sandy and silty soils.
Published literature on research of submarine slopes from the Zeeland Province of
The Netherlands provides more detailed information on the mechanics of failure for
these type of flow slides (Silvas and de Groot 1995; de Groot and Stoutjesdijk 1997;
Stoutjesdijk et al 1998; Koppejan et al 1948). However, the literature from these
sources is limited on specific cases and therefore none of these were included in the
database. The submarine flow slides from Zeeland were generally within loose,
uniformly graded sands and were commonly retrogressive.

3.2.7 FLOW SLIDES IN SENSITIVE CLAYS

The work on “rapid” flow slides from slope failures in sensitive clays is a summary of
the findings from the literature. The focus of much of the research has been on
initiation of landsliding, material properties and the distance of retrogression. Recent
published papers include Locat and Leroueil (1997), Trak and Lacasse (1996), Leroueil
et al (1996), Lefebvre (1996) and Torrence (1996), and earlier publications including
Tavenas (1984), Mitchell and Markell (1974), Bentley and Smalley (1984) and Bjerrum
et al (1969) amongst others. The properties of 15 slides in sensitive clays and data
sources are given in Table 3.8. Further details are given in Section 1 of Appendix B.
The sensitivity, St, of a soil is defined as the ratio of the undisturbed to remoulded
unconfined compressive strength (Equation 3.1). It is often evaluated by comparison of
the undisturbed and remoulded shearing strengths as measured in the field by vane shear
tests. Soils with a sensitivity of greater than about 16 are known as quick clays.

Undisturbed unconfined compressive strength


Sensitivity, St = (3.1)
Remoulded unconfined compressive strength
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.13

Table 3.7: Summary of “rapid” flow slides in natural submarine slopes


Failure Material
Site Trigger Data Sources
Date Description
Orkdalsfjord, 2/5/30 Glacial - loose sand, Exceptional low tide Karlsrud & Edgars (1982),
Norway silt and clay deposits. Edgars and Karlsrud
(1983), Terzaghi (1956)
Bassien Ancient - Earthquake or rapid Edgars and Karlsrud (1983)
sedimentation
Storegga Ancient - Earthquake (?) Edgars and Karlsrud (1983)
Grand Banks 1929 Sand / silt Earthquake Edgars and Karlsrud (1983)
Spanish Sahara Ancient Gravelly clayey sand rapid sedimentation Edgars and Karlsrud (1983)
Walvis Bay (SW Ancient - - Edgars and Karlsrud (1983)
Africa)
Icy Bay / Malaspina Ancient Clayey silt Earthquake Edgars and Karlsrud (1983)
Copper River Ancient Silt / sand rapid sedimentation Edgars and Karlsrud (1983)
Ranger Ancient Clayey and sandy silt rapid sedimentation Edgars and Karlsrud (1983)
(earthquake)
Mid Alb. Bank Ancient silty clay earthquake / rapid Edgars and Karlsrud (1983)
sedimentation
Wil. Canyon Ancient silty clay and silt rapid sedimentation Edgars and Karlsrud (1983)
Kidnappers Ancient sandy silt, clay Earthquake (?) Edgars and Karlsrud (1983)
Kayak Trough Ancient clayey silt rapid sedimentation / Edgars and Karlsrud (1983)
earthquake
Paoanui Ancient silt / sand Earthquake (?) Edgars and Karlsrud (1983)
Mid. Atl. Cont. Ancient silty clay rapid sedimentation Edgars and Karlsrud (1983)
Slope
Magdalena R. 1935 - rapid sedimentation Edgars and Karlsrud (1983)
California Ancient clayey and sandy silt Earthquake (?) Edgars and Karlsrud (1983)
Suva, Fiji 1953 sand Earthquake Edgars and Karlsrud (1983)
Valdez 1964 gravelly silty sand Earthquake Edgars and Karlsrud (1983)
Sokkelvik, Norway 1959 quick clay (?) and - Edgars and Karlsrud (1983)
sand
Sandnessjoen, 1967 - Blasting, man made Edgars and Karlsrud (1983)
Norway fill
Helsinki Harbour, 1936 sand / silt Man made fill Edgars and Karlsrud (1983)
Norway
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.14

Table 3.8: Summary of flow slides in sensitive clays


Site Initial Slide Material Description Data Sources
Classification (Liquefied Material on the
Surface of Rupture)
Description *1 St *2
Lemieuse, Retrogressive Clay (CH), high plasticity, 10 to 100 Lawrence et al (1996)
Ottawa flow slide firm
South Nation Retrogressive Clay (CH), high plasticity, 10 to 100 Lawrence et al (1996)
River, Canada flow slide firm, IL = 1 Eden et al (1971)
St. Jean Vianney, Retrogressive Silty Clay (CL), low > 200 Edgars and Karlsrud (1983),
Quebec flow slide plasticity, cemented, soft Tavenas et al (1971), Trak
to firm, IL = 1.8 and Lacasse (1996), Bentley
and Smalley (1984)
La Grande River, Retrogressive Clayey Silt (CL), low - Lefebvre et al (1991)
Quebec flow slide plasticity, stiff, IL = 1.8
Baastad, Norway Translational Clayey Silt (ML), low 50 to 100 Edgars and Karlsrud (1983)
flow slide plasticity, firm, IL = 1.6 Trak and Lacasse (1996)
Gregerson & Loken (1979)
Rissa, Norway Translational Clay (CL), low plasticity, 100 Edgars and Karlsrud (1983),
flow slide, some very soft, IL = 2.3 Trak and Lacasse (1996),
retrogression Gregersen (1981).

Furre, Norway Translational Silty Clay (CL), low > 70 Hutchinson (1961)
flow slide (some plasticity, firm, IL = 2.1
retrogression)
Vibstad, Norway Translational Clayey Silt (ML), low - Hutchinson (1965)
flow slide plasticity, firm, IL = 1.3
Bekkelaget, Translational Silty Clay / Clayey Silt 80 Eide and Bjerrum (1956)
Norway flow slide (CL/ML), low plasticity,
very soft, IL = 2.4
Tuve, Sweden Retrogressive Silty Clay (CH), high 15 to 40 Duncan et al (1980)
flow slide plasticity, soft, IL = 1.3
Borgen Retrogressive (?) Clay (CL), low plasticity, 20 to 100
flow slide soft, IL = 1.4
Hekseberg Retrogressive (?) Clay, soft to firm 50 to 150
flow slide
Skjelstadmarken Retrogressive (?) Clay (CL), low plasticity, - Edgars and Karlsrud (1983)
flow slide firm, IL = 1.3 Trak and Lacasse (1996)
Verdal Retrogressive (?) Clay (CL), low plasticity, 20 to 200
flow slide soft to firm, IL = 2.5
Selnes Retrogressive (?) Clay (CL), low plasticity, > 100
flow slide soft, IL = 1.9
*1 IL = liquidity index
*2 St = sensitivity, refer Equation 3.1
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.15

3.3 THE M ECHANICS OF SHEARING OF CONTRACTIVE GRANULAR


SOILS
Many “rapid” landslides are the result of a loss of strength of the soil at failure as it
shears in undrained loading under static (monotonic) and cyclic (earthquake) conditions.
This usually occurs in granular soils, which in drained loading would be contractive.
The behaviour of granular soils on shearing in undrained loading is quite complex
and dependent on a number of factors including the material properties, stress
conditions and stress path. This section discusses:
• The general behaviour of granular soils in undrained shear from laboratory testing;
• The identification of soils susceptible to contraction on shearing and development of
liquefaction based on field in-situ tests, particle size distribution and other
characteristics;
• The large strain (residual) undrained strength.

Several aspects of the literature review on soils susceptible to liquefaction and


potential flow are discussed further in Section 2 of Appendix B.

3.3.1 DEFINITIONS

The definition of several key terms used is as follows:


• Liquefaction (or flow liquefaction) – Undrained strain weakening in contractant
soils where, as a result of some disturbance, part (or all) of the load supported by the
soil structure is transferred onto the pore fluid.
• Temporary liquefaction – where a liquefaction condition is only temporarily
developed within a soil mass. The soil mass that displays temporary liquefaction
behaviour is usually contractant on initial shearing (thus liquefaction occurs) but
then dilative as shearing continues (or strain hardening in undrained loading).
• Static liquefaction – the disturbance causing liquefaction is by monotonic loading
and not associated with dynamic events such as earthquake or blasting. Often the
cause of static liquefaction is a change in effective stress conditions such as change
in groundwater level, excavation or rapid sedimentation.
• Cyclic or dynamic liquefaction – liquefaction developed through undrained cyclic
loading in laboratory testing or by earthquake.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.16

3.3.2 CHARACTERISTICS OF CONTRACTIVE AND DILATIVE SOILS SHEARED UNDER


MONOTONIC LOAD CONDITIONS

The observations from monotonic, consolidated undrained triaxial compression tests on


clean sands (Ishihara 1993) indicate that the steady state condition in undrained loading
(Figure 3.2) is dependent on the void ratio alone. For sands at an initial state in e log p'
space (e = void ratio, p' = mean normal effective stress) to the right of the steady state
line, on shearing in drained conditions the sand will contract toward the steady state
line. In undrained loading conditions the tendency for contraction is suppressed and
post peak strain weakening is observed (Figure 3.3) as load is transferred to the pore
fluid and the effective stress path moves to the left (at constant void ratio) toward steady
state.
For sands at an initial state on the left side of the steady state line, when sheared in
drained loading conditions will dilate. In undrained loading conditions the dilation is
suppressed and strain hardening is observed (Figure 3.4) as pore water pressures reduce
and load is transferred onto the soil structure. A “quasi steady state” condition in
undrained loading (where strain hardening follows post-peak strain weakening) is
observed for some initial states as shown in Figure 3.4. It is generally observed where
the initial state is slightly above or slightly below the steady state. Ishihara (1993)
showed the quasi steady state to be dependent on void ratio, initial effective confining
stress and the fabric of the sample, and is therefore not a fundamental state parameter
for sands. The change from strain weakening to strain hardening is termed the phase
transformation.

Figure 3.2: Steady state line of Toyoura sand (Ishihara 1993)


Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.17

Figure 3.3: Undrained behaviour of states on the contractive side of the steady state line
(Ishihara 1993)

Figure 3.4: Undrained behaviour from states close to steady state and on dilative side (a)
stress-strain plot and (b) effective stress paths (Ishihara 1993)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.18

When clean sands are tested in undrained conditions from the same initial void ratio and
different initial effective stress conditions the steady state strength reached for all
samples is similar (Figure 3.5). Yamamuro and Lade (1998) defined three “normal”
states of behaviour for clean sands:
• Stable behaviour, where the undrained stress path is strain hardening. This is
similar to the behaviour observed for clean sands sheared from initial states well to
the left of the steady state line that are strongly dilative and representative of sands
with large negative state parameter, ϕ (see below).
• Temporary instability, where phase transformation or a quasi steady state condition
occurs post the initial peak undrained strength. Decreasing stability is observed
with increasing confining pressure in this region.
• Instability (liquefaction), indicative of the contractive tendency resulting in the
development of positive pore pressures to a steady state condition. Once the
effective stress path exceeds the peak deviatoric stress the sand, in undrained
conditions, becomes unstable and failure occurs. This corresponds to states to the
right of the steady state line (Figure 3.2). Yamamuro and Lade (1998) comment that
this behaviour is observed at relatively high confining pressures, the high confining
pressure suppressing the tendency for dilation.

Figure 3.5: Schematic diagram showing different types of behaviour for undrained
stress paths of samples at the same void ratio and different initial confining pressures
(Yamamuro and Lade 1998).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.19

If we now consider a series of samples tested undrained under monotonic load


conditions from the same initial effective confining stress and different initial void
ratios, the findings from the literature (Ishihara 1993; and others) indicate:
• At very high void ratios, corresponding to sands in a very loose condition, complete
static liquefaction (i.e., zero residual undrained strength) is observed, as typified by
Point c in Figure 3.3. Ishihara (1993) defined the threshold void ratio, e o (refer
Figure 3.2), as the void ratio at or above which samples exhibit zero residual
strength. It corresponds to the void ratio at steady state for zero effective stress.
• At a void ratio slightly less than the threshold void ratio (Points d and e in Figure
3.2) static liquefaction is observed with low residual undrained strength (Figure 3.3).
This corresponds to contractive states located above the steady state line.
• At initial states close to the steady state line temporary instability is observed.
• Stable or strain hardening behaviour is observed at low void ratios, corresponding to
strongly dilative states located well below the steady state line (Figure 3.2).

Two important concepts have been developed to characterise the behaviour of sands.
Been and Jefferies (1985) defined the “state parameter”, ϕ , to uniquely characterise the
state of a sand (Figure 3.6). The state parameter is defined as the difference between the
initial void ratio, e λ , and the void ratio at steady state corresponding to the initial mean

normal effective stress, e ss ( ϕ = e λ − e ss ). The concept of the state parameter is that it


characterises the effective stress path and normalised steady state strength of samples
with identical values of the state parameter, regardless of the initial mean normal
effective stress.
Sladen et al (1985a) proposed the concept of the collapse surface (or instability line)
for evaluation of the liquefaction potential of very loose sands located on the contractive
side of the steady state line. They defined the collapse surface as the locus of the peak
deviatoric stress of undrained triaxial compression tests (Figure 3.7) of samples tested at
the same initial void ratio (but different initial mean normal stress). For liquefaction to
occur the effective stress state of very loose sands must reach the collapse surface in
undrained loading, thereafter in stress controlled tests a static liquefaction condition
develops rapidly. The instability line (or collapse surface) is located at effective stress
states below the steady state. As shown in Figure 3.7, the position of the instability line
and the steady state deviatoric strength for potentially liquefiable sands will vary
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.20

depending on the initial void ratio. It is important to note that under static loading
conditions and in undrained conditions samples are stable at effective stress states below
the instability line.

Figure 3.6: Definition of the state parameter (Been and Jefferies 1985).

Figure 3.7: Collapse surface concept, defined by the locus of the peak deviator stress
from undrained triaxial samples tested at the same void ratio and different initial
effective mean normal stress (Sladen et al 1985a).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.21

The above summary describes the characteristics of sand based on the concept that the
steady state line is a unique parameter for clean sand and that, in undrained loading, the
steady state strength can be determined from void ratio alone. The findings have been
based predominantly on isotropically consolidated, undrained triaxial compressive tests
on clean sands. However, this concept of uniqueness of the steady state line is queried
by the results of research, in particular from work using stress paths other than drained
and undrained compressive testing and also from testing of sands with fines
(Uthayakumar and Vaid 1998; Yamamuro and Lade 1998). This is summarily
discussed below with several aspects discussed further in Section 2 of Appendix B.
For silty sands Lade and Yamamuro (1997) and Yamamuro and Lade (1998) found
that liquefaction was observed at low confining stresses in undrained conditions and
termed this behaviour as “reverse” behaviour to that of clean sands. They considered
that particle re-arrangement of the unstable and highly compressive particle structure
formed due to deposition under very low energy environments as the mechanism for
liquefaction and temporary liquefaction. Under increasing confining pressures the
“normal” states defined for clean sands are observed for silty sands.
Vaid, and his co-workers (Uthayakumar and Vaid 1998; Sivathayalan and Vaid
1998; Vaid and Eliadorani 1998) have shown that the undrained response of loose sand
is highly dependent on the load direction, i.e. whether in triaxial compression or triaxial
extension or simple shear. They showed that samples tested in triaxial extension are
more brittle and mobilize at lower peak strength (i.e. the collapse surface is lower and
the peak undrained strength is lower) than those tested in triaxial compression.
This indicates that the use of triaxial compression tests may not be conservative.
Soils that are dilative under triaxial compression may be contractive under triaxial
extension. Vaid and his co-workers concluded that the ultimate strength is not uniquely
related to void ratio alone, but on the direction of the principal stresses, the effective
mean normal stress prior to shearing, and the magnitude of the intermediate principal
stress.
Imam et al (2002), in researching the yield surface for sands, attributed the
dependency of the undrained response of loose sand on load direction to the inherent
anisotropy associated with deposition of the sand, which existed prior to consolidation.
Higher yielding stresses are observed where the major principal stress is in the direction
of soil deposition. Therefore, for a sand sample formed by vertical deposition the
highest yielding stress in undrained loading is observed for a standard undrained triaxial
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.22

compression test, and the minimum yielding stress for a standard undrained triaxial
extension test.

3.3.3 FIELD METHODS FOR EVALUATION OF FLOW LIQUEFACTION POTENTIAL

Methods for evaluation of liquefaction potential have been developed from the results of
laboratory testing and empirically from field case studies. Most of these methods are
related to cyclically induced liquefaction (i.e., from earthquakes), but it is considered
that much can be learnt from this relative to static liquefaction.
The early work of Harry Bolton Seed pioneered much of the development of
empirical methods from case study analysis for evaluation of liquefaction potential
based on cyclic resistance of the soil under dynamic loading. Robertson and Wride
(1997) discuss the importance of distinguishing between deformations that occur during
cyclic loading (termed cyclic liquefaction) and deformations from flow liquefaction that
usually occurs under the static load after the triggering mechanism (i.e., after an
earthquake). Empirical methods for evaluation of large deformation potential have been
developed for; the Standard Penetration Test (SPT), Cone Penetration Test (CPT), shear
wave velocity from seismic refraction, and the Becker Penetration Test (BPT),
applicable to gravelly soils where the SPT is not appropriate. Empirical correlations to
SPT and CPT are the most widely used given the larger field database of information.
Seed (1979) developed a correlation between the cyclic stress ratio to cause cyclic
liquefaction and the N value of the SPT test based on field data of sand deposits for
essentially level ground condition subject to strong shaking from earthquakes. Seed et
al (1985), recognising that the amount of shear strain observed was dependent on the
relative density of the sand, further developed the correlation to provide guidelines on
the level of shear strain (Figure 3.8) with respect to SPT and cyclic resistance ratio for
clean sands. Robertson and Wride (1997) comment that much larger shear strains can
be expected for loose sands ((N1)60 < 15), where effective stresses may reduce to zero
under dynamic loading, than for denser sands ((N1)60 > 15) where large pore pressures
may develop but effective stresses ma y not fully reduce to zero. The line on Figure 3.8
for shear strains of 20% could be considered as representing a boundary of flow
liquefaction conditions.
Baziar and Dobry (1995) provide an indication of flow liquefaction potential for
silty sands and sandy silts (≥ 10% fines) based on case histories of flow failure and large
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.23

ground deformation all triggered by earthquakes corresponding to a moment magnitude


less than or equal to 8. They hypothesise that the right boundary in Figure 3.9 is the
boundary between flow liquefaction and no-flow conditions (no-flow for (N1)60 to the
right of this boundary).

NCEER = National
Centre for Earthquake
Engineering Research

Figure 3.8: Correlation for cyclic resistance ratio (CRR) based on corrected SPT blow
count, (N1)60, for clean sands under level ground conditions (Robertson and Wride
1997)

Ishihara (1993) compared the threshold flow liquefaction boundary determined from
laboratory undrained monotonic compression tests on clean sands and silty sands
(separate comparison for sands and silty sands) to known cases of field flow
liquefaction triggered by earthquake (Niigata in 1964 and Nihonkai-Chuba in 1983)
where SPT N value data was known prior to the earthquake. The laboratory sand and
silty sand samples were not from the field sites, but were considered by Ishihara as
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.24

typical of the liquefied sands. The flow liquefaction boundary for the laboratory
samples was defined from the quasi-steady state for water sedimented samples. This
boundary was converted into a plot of SPT N value as a function of effective
overburden stress, using an empirical correlation derived from the correlation between
density ratio and N value (Equation 3.2) suggested by Skempton (1986). Ishihara
(1993) found the flow liquefaction no-flow boundaries (in SPT N value versus effective
over-burden pressure space) of the laboratory samples to be relatively diverse between
samples and commented on the marked difference in density ratio at which the quasi-
steady state occurs for clean sands. However, the SPT N values were in the same range
as the field values from actual sites of flow failure.

N1
=a+b (3.2)
(D r 100 )
2

where N1 is the SPT N value corrected for over-burden pressure, Dr the relative density
of the sand, and a and b constants based on the fines content and mean particle diameter
of sand particles.

Figure 3.9: Corrected SPT blow count, (N1)60, versus vertical effective over-burden
′ ) for saturated non-gravelly silty sand deposits that have experienced large
pressure ( σ vo
deformations and were triggered by earthquake (Baziar and Dobry 1995)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.25

Figure 3.10a presents the boundary between flow liquefaction and no-flow, and the
boundary between cyclic liquefaction and no-liquefaction estimated by Ishihara (1993)
for sands with fines contents up to 30%. The flow liquefaction no-flow boundary
represents the two boundary curves (one from clean sands and one from silty sands)
giving the maximum possible N value determined from the comparisons between
laboratory testing and field values from actual sites of flow failure, and can therefore be
considered conservative. Figure 3.10b is the boundary between flow liquefaction and
no-flow in terms of qc, the CPT tip resistance, determined by conversion of the SPT N
value boundary using a correlation between qc and N. This boundary is compared to the
boundaries proposed by Sladen and Hewitt (1989) and Robertson et al (1992).

Figure 3.10: Flow liquefaction and no-flow bounds in terms of (a) field measured SPT
blow count and (b) CPT qc value from flow liquefaction case histories (static and
earthquake triggered) and laboratory undrained testing (Ishihara 1993).

More recently, Cubrinovski and Ishihara (2000) have researched the effects of particle
size distribution, fines content and particle shape on the flow liquefaction potential of
sands and silty sands (with up to 30% non-plastic fines content). From analysis of
published laboratory results for some 52 sandy soils they found that the flow
liquefaction potential (and steady state line) could be defined based on the value of void
ratio range (emax – emin) and consideration of particle shape. They found that at a given
effective mean normal stress the relative density (or density index) defining the flow
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.26

liquefaction no flow boundary increased with increasing void ratio range. Figure 3.11
presents the comparison of flow liquefaction potential (Cubrinovski and Ishihara 2000)
for the void ratio ranges 0.35 and 0.60 for both round-grained sands and angular grains.
The shaded area above the line (in density ratio to mean normal stress space) defines the
zone of flow liquefaction potential. For a void ratio range of 0.60 the relative density, at
a given mean normal stress, below which the sand is susceptible to flow liquefaction is
much greater than for a sand with a void ratio range of 0.35.

Figure 3.11: Influence of void ratio range on flow potential of sandy soils for (a) round-
grained sands and (b) angular grains (Cubrinovski and Ishihara 2000)

Typical values of void ratio range (emax – emin) from their data set ranged from 0.25 to
0.65 for sands with up to 30% non-plastic fines content. For a given grain shape, the
factors affecting void ratio range were fines content (the greater the fines content the
greater the void ratio range) and mean grain size, D50 (the greater D50 the lower the void
ratio range). Therefore, the findings indicate that finer sands and silty sands are more
susceptible to flow liquefaction than coarser and cleaner sands.
Cubrinovski and Ishihara (2000), from their database, derived empirical correlations
of relative density (at the threshold void ratio) and slope of the steady state line, λ , with
respect to the void ratio range. As shown in Figure 3.12 and Figure 3.13, whilst the data
correlate reasonably well with the trendlines, there is a significant scatter. Cubrinovski
and Ishihara (2000) then, using their derived correlation of relative density to SPT N
value corrected for over-burden pressure, N1, (Figure 3.14), derived empirical
relationships to define the flow liquefaction no-flow boundary in terms of field
measured SPT blow count. These equations (Equation 3.3 for round-grained sands and
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.27

Equation 3.4 for angular sands), by Cubrinovski and Ishihara (2000), define the field
measured SPT N value corresponding to the flow liquefaction no-flow boundary, Ns, in
terms of effective overburden stress, σ ′v , void ratio range (emax – emin) and mean normal
stress, p′ .

0.01 − (1 − log p ′){− 0 .01 + 0 .125(e max − e min )}


2
 σ ′v 
0. 5
9 
Ns =    − 0.4 + 1 .4(e max − e min ) + 
(e max − e min ) 1.7  98   (e max − e min ) 
(3.3)

0. 01 − (1 − log p ′){− 0.02 + 0 .25(e max − e min )}


2
 σ ′v 
0. 5
9 
Ns =    − 0. 4 + 1.4 (e max − e min ) + 
(e max − e min ) 1.7  98   (e max − e min ) 
(3.4)

Figure 3.12: Relationship between relative density, Dro , at the threshold void ratio, eo,
and void ratio range (emax – emin) for sandy soils (Cubrinovski and Ishihara 2000).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.28

Figure 3.13: Relationship between the slope of the steady state line, λ , and void ratio
range for (a) round-grained sands and (b) angular grained sands (Cubrinovski and
Ishihara 2000).

2
Figure 3.14: Correlation between N1 Dr (N1 = SPT blow count corrected for over-
burden pressure, Dr = relative density) and void ratio range for cohesionless soils
(Cubrinovski and Ishihara 2000).

Graphical representations of the flow liquefaction no-flow boundary for both round-
grained and angular sands according to Equations 3.3 and 3.4 are shown in Figure 3.15
for void ratio range from 0.3 to 0.6 and lateral stress ratios, Ko, of 0.5 and 1.0.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.29

It is important to note that the National Standards for determining the minimum void
ratio are not consistent between countries and this will therefore affect the void ratio
range and relative density (density index). The Australian Standard is likely to result in
a lower minimum void ratio than the Japanese Standards referred to by Cubrinovski and
Ishihara (2000). In addition, the scatter associated with derivation of the empirical
correlations should also be considered when applying these relationships.

Figure 3.15: Field measured SPT blow count boundaries differentiating flow
liquefaction no-flow conditions for void ratio range of 0.3 to 0.6 for (a) round-grained
sands and (b) angular sands (Cubrinovski and Ishihara 2000)

Cubrinovski and Ishihara (2000) also defined the region of flow liquefaction at zero
residual undrained strength (i.e., for void ratios at or above the threshold void ratio,
defined in Section 3.3.2) in terms of SPT N value, using Equations 3.3 and 3.4. From
Figure 3.11 the region of flow liquefaction at zero residual undrained strength is at
relative densities less than at the intercept of the liquefaction flow no-flow boundary;
i.e., at zero effective mean normal stress. For example, from Figure 3.11a flow
liquefaction at zero residual strength would be at relative densities less than about 10%
for a void ratio range of 0.35 and less than about 43% for a void ratio range of 0.6 (both
for round-grained sands). Figure 3.16 is an example showing the liquefaction flow no-
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.30

flow boundary and the region of flow liquefaction at zero residual undrained strength in
terms of SPT N value and effective over-burden pressure.

Figure 3.16: Flow characterisation of sand based on field measured SPT blow count
showing liquefaction flow no-flow boundary, and region of flow at zero residual
undrained strength (Cubrinovski and Ishihara 2000)

3.3.4 EFFECTS OF NON-PLASTIC FINES AND GRAVELS BASED ON FLOW


LIQUEFACTION POTENTIAL

Byrne and Beaty (1998) considered the question of whether fines in sands were an
instigator or inhibitor of liquefaction potential, and commented that the structure of the
soil had a significant influence on the characteristics of sands with fines. Robertson and
Wride (1997) commented that from field case studies it was not clear what effect fines
had, whether the cyclic liquefaction resistance of the soil is increased or whether the
(N1)60 value of the soil decreases as a consequence of the increase in compressibility and
decrease in permeability associated with increasing fines content.
The results from undrained monotonic laboratory testing on silty sands with non-
plastic fines (Cubrinovski and Ishihara 2000; Lade and Yamamuro 1997) indicate that,
in terms of relative density, the fines content significantly increases the flow
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.31

liquefaction potential of sands as indicated by Figure 3.11 and Figure 3.17. Lade and
Yamamuro (1997) comment that the fines have a significant role in the particle
structure, resulting in highly compressible soils with increased liquefaction potential.
They also recognised the findings of field-testing correlations with actual earthquake
induced liquefaction events that indicated sands with significant fines content were
more resistant to cyclic liquefaction. Lade and Yamamuro (1997) considered that the
effects of stress history and aging (creep or cementation) are likely to increase the field
resistance of silty sands. However, for newly constructed hydraulic fills and very recent
natural deposits they considered that fines content would decrease the liquefaction
resistance.
Field evidence (Eckersley 1985; Dawson et al 1998; Bishop 1973; Siddle et al 1996)
indicates that sandy gravels, typically with low fines content, are susceptible to static
liquefaction and flow sliding. This is supported by observations of static liquefaction in
laboratory undrained monotonic tests on materials similar to those for which flow
liquefaction occurred in the field (Dawson et at 1998; Dawson 1994; Eckersley 1986).
The results from Dawson (1994) on the Fording mine over-burden dump sandy gravel
material are presented in Figure 3.18 and show that at low confining pressures the
sample was strain-weakening on shearing in undrained loading.
Cubrinovski and Ishihara (2000) comment that coarse sands and gravels are
practically safe against flow liquefaction due to the typically narrow void ratio range of
these soil types.

Figure 3.17: Increase in liquefaction potential (as measured by relative density) with
increasing content of non-plastic fines (Lade and Yamamuro 1997)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.32

Figure 3.18: Fording sandy gravel sample (a) undrained triaxial compression test results
and (b) consolidation and steady states (Dawson et al 1998).

3.3.5 RESIDUAL UNDRAINED SHEAR STRENGTH OF FLOW LIQUEFIED SOILS

Methods developed for assessment of the residual undrained shear strength of flow
liquefied soils include empirical correlations based on back analysis of mostly
earthquake induced liquefaction flow failures (Stark and Mesri 1992; Seed 1987; Seed
and Harder 1990; Baziar and Dobry 1995), and estimation from laboratory test results.
Whilst undrained conditions can be guaranteed in the laboratory and the term residual
undrained strength is appropriate at large strains, in the field cases some drainage may
occur. However, the term residual undrained strength is used here for both the
laboratory and field cases.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.33

Prediction of the residual undrained shear strength from laboratory test results
would appear to be a relatively straightforward process. As shown in Figure 3.7, the
residual undrained shear strength can be estimated from knowledge of the field void
ratio and the steady state line. However, it is not this simple. Seed (1987) reported that
experience with the laboratory method seemed to give values of residual undrained
shear strength much greater than those determined by back analysis of field case studies.
Seed queried the use of laboratory methods to field situations with respect to the effects
of water re-distribution and the assumption of constant void ratio, particularly for
stratified soil deposits and for liquefaction triggered by earthquake. At the National
Science Foundation workshop on the shear strength of liquefied soils (NSF 1998) the
significant debate amongst eminent researchers in this field over aspects of the
theoretical fundamentals is indicative of the limitations of knowledge in this area.
As discussed in Section 3.3.2, it is the void ratio and effective mean normal stress
that determine the behaviour of cohesionless sands and silts (and gravels to some
extent) in undrained conditions. At void ratios less than the threshold void ratio, the
residual undrained shear strength (for contractant samples) is determined from the void
ratio (Figure 3.2 or Figure 3.7) and will increase with decreasing void ratio. At void
ratios greater than or equal to the threshold void ratio (Figure 3.2) the laboratory
residual undrained shear strength is zero. The recent research by Cubrinovski and
Ishihara (2000) indicates that the relative density defining the threshold void ratio is
greatly dependent on the void ratio range (Figure 3.11). For silty sands (typically broad
void ratio range) zero residual undrained shear strength is observed in laboratory
undrained triaxial tests at relative densities as high as 40 to 50%. For clean, medium to
coarse grained sands (typically narrow void ratio range) zero residual undrained shear
strength is generally observed at much lower relative densities, typically less than about
10%.
Seed (1987) back analysed some 12 case studies of flow sliding (or large
deformation) and developed a correlation of residual undrained strength to the
equivalent clean sand SPT blow count, (N1)60-CS. He commented that this would
provide a useful guide to estimation of the residual undrained strength of flow liquefied
soils. Figure 3.19 presents the correlation from Seed and Harder (1990) whom added to
the Seed (1987) case studies and also considered dynamic effects in the travel of the
flow slide. They recommended use of the lower bound relationship for estimation of
the residual undrained strength.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.34

Figure 3.19: Relationship between equivalent clean sand SPT blow count, (N1)60-CS, and
residual undrained strength from field case studies (Seed and Harder 1990)

Stark and Mesri (1992) proposed a correlation (Equation 3.5) between the residual
undrained shear strength ( S u (crit ) ), “normalised” with respect to the initial effective

′ ), and the equivalent clean sand SPT blow count, (N1)60-CS.


over-burden pressure (σ vo

They used the yield strength correction factor for fines content (Table 3.9) whilst Seed
(1987) and Seed and Harder (1990) used the residual strength correction factor.
Equation 3.5 represents a lower bound through the laboratory-undrained test data used
to establish the correlation and is shown in Figure 3.20. Flow slides in silty sands and
sandy silts plot close to the correlation and Stark and Mesri comment that this indicates
limited drainage during flow. The case studies in sands and gravelly sands with limited
fines contents plot well above the line indicating significant drainage during flow.

′ = 0.0055 × ( N 1 )60− CS
S u (crit ) σ vo (3.5)

Baziar and Dobry (1995) further analysed the case histories used by Seed and Harder
(1990) and Stark and Mesri (1992). They considered those case histories of flow slides
and large lateral spreads in silty sand deposits (greater than 10% fines) triggered by
earthquake and concluded that the residual undrained shear strength (Sr) from back
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.35

analysis was related to the initial effective over-burden pressure, with the limits of
S r σ ′vo in the range 0.04 to 0.2 (Figure 3.21).

Figure 3.20: Relationship between undrained critical strength ratio and equivalent clean
sand SPT blow count, (N1)60-CS (Stark and Mesri 1992)

Table 3.9: SPT blow count ((N1)60) yield strength and residual strength correction
factors for fines content (Seed et al 1985; Seed 1987)
Fines Content (N1)60 Correction Factor (N1)60 Correction Factor
(%) for Yield Strength for Residual Strength
10 2.5 1
15 4 -
20 5 -
25 6 2
30 6.5 -
35 7 -
50 7 4
75 7 5

Riemer (1998) comments that there is little theoretical basis for correlation of the
residual undrained shear strength to effective vertical stress. Poulos (1998) comments,
and the observations of undrained laboratory tests indicate, that the residual undrained
shear strength is a function of void ratio only. Poulos (1998) and Martin (1998) indicate
that some correlation of residual undrained strength to the initial effective over-burden
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.36

pressure may be observed due to consolidation effects and increased relative density
under increasing vertical stress; however, the ratio would not be constant and would
vary for different soil types.

Figure 3.21: Residual undrained shear strength, Sr, to vertical effective over-burden
pressure for silty sand deposits (≥ 10% fines), triggered by earthquake, that have
experienced large deformation (Baziar and Dobry 1995)

In concluding, it is considered that there is significant uncertainty associated with


estimation of the residual undrained shear strength, both from laboratory testing
methods and from correlations with case histories, and any method should be used with
caution.
With respect to laboratory testing, there is an element of uncertainty with respect to
fundamental theoretical issues and these issues significantly affect determination of the
residual shear strength of the liquefied soil. Although the methodology for assessment
based on laboratory testing is relatively straight forward, application of theoretical
concepts from laboratory testing have proven difficult to apply to site conditions, the
difficulty being associated with obtaining “undisturbed” samples of cohesionless soils,
the sensitivity of the soil behaviour to small changes in void ratio, complexity of factors
such as in-situ stratification and the actual test results themselves. No consensus was
reached at the National Science Foundation Workshop on the shear strength of liquefied
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.37

soils (NSF 1998) regarding sample preparation, the type of laboratory testing to be
undertaken, and the evaluation of the residual undrained strength from the results.
Most of the empirical methods derived from back analysis of historical case studies
have been from flow slides and large deformations triggered by earthquake; although,
the manner in which flow liquefaction is triggered (either static or cyclic) will not affect
the large strain residual undrained strength. One of the difficulties with using empirical
correlations based on SPT blow count data is considered to be that it does not (and
cannot) incorporate all the material properties that affect flow liquefaction behaviour.
Of the available empirical methods, it is considered that the lower limit of the Seed
and Harder (1990) correlation (Figure 3.19) provides a reasonable estimate of the
residual undrained shear strength. The shape of the relationship is similar to that which
would be obtained between residual undrained strength and relative density (or void
ratio) from laboratory results, with close to zero residual undrained strength at low
relative densities, and increasing relatively rapidly with increasing relative density. The
Stark and Mesri (1992) correlation is also considered to provide reasonable estimates of
residual undrained strength. Direct correlations of residual undrained strength with
effective over-burden stress, such as by Baziar and Dobry (1995), are considered
questionable as they do not consider the fundamental relationship between residual
undrained shear strength and void ratio. The results of back-analysis indicate that
drainage during flow sliding is an important aspect in travel behaviour and the potential
for drainage during flow must be considered (Stark and Mesri 1992; Castro 1998). For
the materials of lower permeability (e.g., silty sands) back-analysis results indicate that
little drainage during flow sliding occurs compared to the cleaner sands.
Figure 3.22 presents guidelines for assessment of the residual undrained shear
strength of liquefied soil that came out of the 1997 National Science Foundation
Workshop on the shear strength of liquefied soil (NSF 1998). The working group
recommended the effort and complexity of field and laboratory testing for site
assessment be dependent upon project risk. They devised three risk categories; low,
moderate and high, for which they recommended assessment comprise:
• For “Low Risk” projects, site investigation consisting of cone penetration testing
(CPT) or piezocone testing (CPTU) in combination with limited SPT testing,
followed by limited classification type laboratory testing. Estimation of residual
undrained shear strength using existing empirical correlations based on historical
cases.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.38

• For “Medium Risk” projects - inclusion of additional in-situ testing, such as


geophysical methods, and inclusion of laboratory testing on reconstituted samples
for assessment of the critical state line. Estimation of residual undrained shear
strength using existing empirical correlations and observations of soil behaviour
from the laboratory testing.
• For “High Risk” projects a two-phase investigation is recommended. Identification
of potential critical areas from the initial phase based on predicted behaviour using
estimates from empirical correlations. Detailed second phase investigation aimed at
reducing uncertainty from the preliminary phase. Laboratory testing on undisturbed
samples (taken using ground freezing techniques) in the detailed stage that could
include triaxial compression/extension, simple shear, and/or torsional ring shear
tests. Estimation of residual undrained shear strength from results of the detailed
laboratory testing program.

Figure 3.22: Flow chart for assessment of residual undrained shear strength of liquefied
soil (NSF 1998)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.39

3.4 M ECHANICS OF F AILURE OF F LOW SLIDES


This section discusses the mechanics of failure of flow slides in contractile soils.
Within each slide sub-group the properties of the slide material and factors influencing
the triggering of and development of slope failure are drawn from the literature and case
study data.
It is important to distinguish between the trigger mechanism/s for the processes
leading up to slope instability and the occurrence of “rapid” sliding. Once an unstable
slope condition is reached the progression to “rapid” sliding is dependent on the slope
geometry and material properties within the unstable slope. For example, if a static
liquefaction condition is triggered in part of a slope a “rapid” flow slide event is not
necessarily the outcome.

3.4.1 MECHANICS OF DEVELOPMENT OF FLOW SLIDING

The characteristic common to all soils susceptible to static liquefaction is that the soil
structure is prone to contract in drained shear. In undrained loading, a loss in undrained
strength occurs as a portion of the load once supported by the soil structure is
transferred onto the pore fluid. In most cases the soil is in a saturated condition, but
laboratory testing (Vaid and Eliadorani 1998) indicates that liquefaction can be
triggered in near saturated states.
The mechanism of initiation of flow sliding under static conditions is summarised as
follows:
• The effective stress state of the saturated (or near saturated) soil within which
liquefaction is initially triggered is at or above the instability line or collapse surface
(Figure 3.23). It is possible for soils to exist in drained conditions above the
instability line even though they are contractive on shearing in drained conditions
under these stress conditions (see below).
• A change in effective stress conditions is required to trigger liquefaction for
effective stress states existing above the instability line (i.e. within highly stressed
regions of the slope). For effective stress states below the instability line a change
in effective stress conditions is required to move the effective stress state to the
instability line at which point a liquefaction condition can be triggered.
• Deformation occurs as a result of the change in effective stress conditions.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.40

• At effective stress states at or above the instability line the soil structure is
contractile on shearing and it is within these highly stressed regions of the slope that
a localised static liquefaction condition is initially developed. As a result of this
propensity for contraction on shearing, load once supported by the soil structure is
transferred onto the pore fluid. Undrained conditions will persist if the resulting
pore water pressures cannot be dissipated.
• Once liquefaction has initiated the load once supported by the soil in the localised
liquefied zone is transferred onto the soil structure in the immediate vicinity.
Progression of the zone of liquefaction will occur if the adjacent soil structure
cannot support the additional stresses imposed. The rate of progression is dependent
on the loss in undrained strength of the soil on liquefaction and the additional load
capacity of the adjacent soils. The undrained brittleness index (Figure 3.24), IB, is a
measure of the loss in undrained strength. The higher the value of IB the more likely
the faster the rate of progression of the liquefied zone will be.
• Failure of the slope will occur if or when it has reached an unstable condition.
• The kinetic energy of the failed mass is, to some extent, dependent on the undrained
brittleness index. High kinetic energy is associated with high undrained brittleness
index.

Deviator Critical
Stress, q State Line

Instability line (or


collapse surface)
Steady state
condition
Initial stress state

Development of undrained
conditions and liquefaction

Mean Effective Stress, p´

Figure 3.23: Idealised mechanism for initiation of flow slides.


Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.41

Figure 3.24: Definition of undrained brittleness index, IB, in strain weakening soil.

As discussed above, a change in the effective stress state is the trigger for static
liquefaction and potential flow sliding. The forms of trigger and effective stress
changes are considered to include:
• Triggering of static liquefaction at effective stress states at or above the instability
line. Lade (1992) considered that the states of stress on some planes in the soil mass
within the slope are in the region of potential instability, i.e., above the instability
line (Figure 3.25 as an example of a submarine slope). In drained conditions the
slope is in a stable condition; however, Lade (1992) considers that a small
disturbance, in soils that have a relatively low permeability will trigger static
liquefaction that could result in development of a flow slide. Flow slides in
submarine slopes triggered by large tidal changes are an example. This mechanism
could also be appropriate to the initial failure for flow slides in tailing dams and
sensitive clays.
• Rising groundwater levels that move the effective stress state to the instability line.
Dawson (1994) proposed this mechanism (Figure 3.26) for flow slides in loose
dumped coal mine waste spoil piles in British Columbia, Canada. On reaching the
instability line static liquefaction occurs within critically stressed regions of the
slope resulting in the development of an unstable slope and triggering a flow slide.
• Reduction in soil suction of partially saturated soils due to water infiltration and
wetting up of the soil mass. Flow slides in loose fills in Hong Kong are considered
to fail by this mechanism. The effect of decreasing soil suction is to push the
effective stress state toward the instability line.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.42

• Excavation, erosion and rapid sedimentation resulting in over-steepening and


heightening of slopes are considered to move the effective stress state toward the
instability line. Leroueil et al (1996) consider these factors to be predominantly
aggravating factors, but they can also trigger a slope failure.
• Layering in the soil profile is considered to have an important influence.
Differences in soil stiffness due to layering will result in concentration of shear
strain at the interface and can trigger liquefaction in this region. Examples of this
are evident from the analysis of failures of embankments on soft ground (Chapter 4)
and it is considered a significant factor for a number of failures across the broad
spectrum of flow slides considered here.

Figure 3.25: Effective stress states on some planes (submarine slopes) located within the
region of potentially instability (Lade 1992).

A static liquefaction condition will only be triggered where pore water pressures
developed due to contraction on shearing cannot be dissipated. The factors, some of
them inter-related, that are considered to influence the development of undrained
conditions and therefore flow sliding in contractive soil structures include; soil
permeability, the rate of shearing, amount of potential volume contraction, effective
stress level and the undrained brittleness index. The effect of soil permeability and rate
of shearing are considered to be as follows:
• For fine-grained soils (clays, silts, sandy silts, etc.) of low permeability, low rates of
shear strain are required to trigger static liquefaction. This is evident from the
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.43

limited pre-failure deformation behaviour observed for flow slides in these soil
types.
• For the coarser grained, more permeable soil types, such as sandy gravels and
gravelly sands, relatively high strain rates are required to trigger static liquefaction.
This is evident from the pre-collapse strain rates of up to 15 to 50 m/day and
deformations of at least several metres for flow slides in coal waste spoil piles in
Canada and South Wales (refer Section 3.7.1.1).

Figure 3.26: Collapse model for mine waste dumps (Dawson et al 1998)

3.4.2 FLOW SLIDES IN GRANULAR STOCKPILES

(a) Material Properties


The particle size distribution and relative density of the stockpiled granular material are
considered the two most significant material properties influencing the likelihood or
otherwise of susceptibility to static liquefaction.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.44

Figure 3.27 presents the approximate particle size distributions for coal waste and
coking coal susceptible to static liquefaction and within which flow slides have
developed in British Columbia, South Wales and Hay Point. The materials classify as
sandy gravels with some to a trace of silt sized fines (1.5 to 12% finer than 75 micron).

100

1
80
4
Percent passing (%)

60
2
3

40

20

0
0.01 0.1 1 10 100
Silt Sand Gravel
Particle size (mm)
1 - Aberfan, Tip 7 (Bishop et al 1969)
2 - Aberfan, Tip 7 (Hutchinson 1986)
3 - Coal mine waste spoil piles, South Wales. One standard deviation around the mean (Taylor 1984).
4 - Coal mine waste spoil piles, British Columbia - materials susceptible to liquefaction (Dawson et al 1998)
5 - Coking coal stockpiles, Hay Point (Eckersley 1990)

Figure 3.27: Particle size distribution of coal mine waste and coking coal susceptible to
static liquefaction

It is considered from the available case studies that the coarser side of the grading
curves in Figure 3.27 are close to representative of the upper bound of materials within
which liquefaction and flow sliding can occur. Particle size distributions coarser than
this, particularly at the finer end of the grading curve, have sufficient permeability that
pore water pressures developed due to contraction of the structure on shearing can be
dissipated. Flow slides have not developed in coarser materials such as the coarser
sections of the spoil piles in British Columbia or within dumped dirty rockfills used in
construction of dam embankments that have undergone significant settlements on first
filling and drawdown.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.45

Density is a significant factor in susceptibility of a material to static liquefaction.


Laboratory undrained tests on coal waste and coking coal (Section 3.3.4) indicate that at
void ratios greater than 0.3 the sandy gravels typical of the stockpiles in British
Columbia and Hay Point are susceptible to static liquefaction, and that the undrained
brittleness increases is with increasing void ratio above 0.3. Field testing indicates that
void ratios in excess of 0.3 are obtained for these materials in a loose dumped condition,
but it is dependent on the method of placement and to some extent moisture content at
placement. For the flow slides at Hay Point (Eckersley 1985, 1986) most failures
occurred within bulldozed stockpiles, for which Eckersley (1985) comments is a result
of the low dry densities achieved for this method of placement (compared to placement
from the aerial stackers). The density profile of the stockpiled coal waste of Tip 7 at
Aberfan (from Bishop et al (1969) data) indicates densities as low as 82% to 91% of the
Standard Proctor Maximum Dry Density at depths up to 15 m to 20 m (equivalent to
void ratios greater than about 0.43), indicative of the loose state of the coal waste placed
by tippler.
For the coal mine waste spoil piles at British Columbia and South Wales tipping is
(or was) from a low height onto the active portion of the tip. It is notable from the
literature that no “rapid” flow slides are reported to have developed from failures in coal
mine waste spoil piles formed by dumping from draglines. The findings would indicate
that the stockpiling technique affects the material density such that likelihood of flow
sliding is negligible in coal waste placed by dragline and much less likely in coal placed
by stackers.
Flow slides have been reported in stockpiles of other than coal waste and coking
coal. Su and Miller (1995) report flow slides, from China, in waste spoil piles from the
mining of iron ore, copper, coal and non-ferrous metals. They indicate that the spoil
piles contain a proportion of fine granular material derived mainly from weathered rock
and/or overburden soils and that the materials are loosely placed by dumping at the crest
of the advancing face of the spoil pile. Bishop (1973) reports on flow slides in loose
spoil piles of well graded sands (from China clay mining) and flyash.

(b) Factors Influencing Flow Sliding in Granular Stockpiles


From analysis of the case studies and review of published literature it is evident that
there are several key factors that influence the likelihood of a flow slide occurring:
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.46

• Particle size distribution and material density (or placement technique) as discussed
above.
• Rainfall and snow/ice melt. Analysis of the case studies indicates that rainfall is a
significant triggering factor for flow slides in South Wales and Hay Point, and both
rainfall and snow or ice melt are significant factors for flow slides in British
Columbia. Su and Millar (1995) indicate that heavy rainfall is a significant factor
for the triggering of flow slides in waste spoil piles in China. The case study data
indicates:
− For British Columbia more than 60% of flow slides occurred in spring and early
summer, the time of snowmelt. The other significant period for flow sliding in
autumn, generally the wettest period.
− For South Wales all flow slides (where the date of the landslide is known)
occurred between October and February, the wettest months in South Wales. A
significant, although not severe, period of rainfall (typically 1 in 2 to 1 in 5 year
event) usually preceded flow sliding.
− For Hay Point flow sliding was associated with significantly wet periods.
• Foundation slope angle for slides in coal waste spoil piles. For British Columbia
Dawson et al (1998) report that flow slides tended to occur where foundation slope
angles exceeded 15 degrees. Golder Assoc. (1994) report the average foundation
slope of 46 flow slide events in British Columbia to be 25.4 degrees (standard
deviation of 6.4 degrees). For South Wales the foundation slope angles of flow slide
events ranged from 8 to 27 degrees (average of 17.2 degrees and standard deviation
of 5 degrees). For the flow slides in spoil piles in China, although not directly
discussed by Su and Millar (1995), they imply steep foundation slopes were present
in a number of the failures. The China clay spoil piles were constructed on
foundations sloping at 6 and 7 degrees and the flyash failure in Jupille on a hillside
(Bishop 1973).
• Active dumping is usually in progress leading up to the failure (but not always).
• Springs, streams or high pore pressures in the foundation. In 55% of the case
studies from South Wales the toe of the spoil pile encroached over an existing spring
or stream. A significant number of the tips in South Wales were constructed over
the Pennant Sandstone formation (massive sandstone with high permeability along
discontinuities) and the springs were generally associated with the outcrop of inter-
bedded argillaceous seams, coal seams or surficial glacial boulder clay deposits.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.47

• Construction of the tip over pre-existing earthflows (specific to South Wales).


Almost 44% of the flow slide case studies from South Wales were within tips
known to have been constructed over pre-existing earthflows.

Siddle et al (1996) observed that within the United Kingdom the coal waste spoil
piles at South Wales are the only geographic region (in the UK) where flow sliding of
these materials is observed, the reasons being the significantly wetter climate and
steeper topography. From Taylor (1984) it is evident that for the South Wales spoil
piles the material grading was generally on the coarser side of average for coal spoil
piles in the UK, whilst the method of placement (in the early to mid 1900’s) did not
vary significantly, indicating that grading and placement method are not significant
factors for the confinement of flow sliding in the UK to South Wales.
The foundation conditions and hydrogeology are also considered significant factors
in the occurrence of flow slides at South Wales. At Aberfan, piezometers installed into
the Pennant Sandstone foundation indicated a significant rise in pore pressure following
heavy rainfall (Bishop et al 1969) and an increase in the foundation pore pressures in
the toe region of the tip, and this may have been the trigger for initial instability in the
toe region of the spoil pile which encroached over an existing spring.
Saturation or near saturation of the coal waste or coking coal (or other liquefaction
susceptible loose stockpiles granular material) is a requisite for static liquefaction. It is
possible that melting of ice and snow within the coal waste spoil piles at British
Columbia is sufficient to saturate the finer sandy gravel sized mine spoil within which
static liquefaction can occur. At South Wales and Hay Point saturation of the lower
portion of the waste is likely following heavy rainfall due to the lower permeability of
the foundation. At Aberfan, piezometers located close to the base of the tip recorded
excess pore pressures of up to 3 m following heavy rainfall and maintained this level for
the subsequent 1 to 2 days (Bishop et al 1969).
High rates of shearing seem to be a requisite for triggering of undrained conditions
and therefore potential flow sliding in these coarse grained materials as indicated by the
pre-failure deformation observations (refer Section 3.7.1.1). In addition, the pre-
collapse shearing may also be creating the critical stress and void ratio conditions
needed for development of static liquefaction.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.48

(c) Failure Mechanism for Flow Sliding in Granular Stockpiles


The specific mechanics of failure for flow slides in granular stockpiles from analysis of
the case studies and published literature is not clear. The possible mechanisms
discussed in Section 3.4.1 are the collapse model proposed by Dawson (1994) and
Dawson et al (1998) for mine waste failures in British Columbia (Figure 3.26) and the
model proposed by Lade (1992), Figure 3.25.
The Lade (1992) model is considered to have merit given the likely high stress
conditions present within the steep spoil and stockpiles. Together with the low density
of the materials, the stress conditions under which contractile conditions on shearing are
prevalent is considered likely. Shearing along interfaces that are located within the
instability region (e.g. between the spoil and foundation or between the fine and coarse
zones for British Columbian waste stockpiles) as a result of changes in effective stress
conditions (loading for active tips or changes in pore pressure conditions) could then
trigger static liquefaction and result in failure and flow sliding. High stress conditions
are also considered to develop in the toe region of the spoil piles due to stress re-
distribution from deformation of the spoil pile and changes in foundation slope.
Whilst this mechanism may not be solely responsible for triggering of static
liquefaction, leading to slope instability and flow sliding, it is considered significant in
the overall mechanism. Bulging observed in the lower portion of the spoil pile prior to
failure and the results of numerical modelling (Dawson 1994) are considered to support
the likelihood of high stress conditions in the toe region. In addition, the influence of
foundation slope for flow slides in British Columbia and South Wales would tend to
support the potential for development of shear strains in saturated to near-saturated
regions of the slope where contractile conditions are prevalent.
Hungr and Kent (1995), commenting on the findings by Golder Assoc. (1994),
indicate that static liquefaction could be triggered within the foundation of the spoil pile
failures in British Columbia. They consider that the upper zone of the foundation could
potentially be susceptible to static liquefaction due to weakening from frost heave
processes. Dawson et al (1998) also comment that other processes can act as a collapse
trigger such as shearing in the toe region, creep and weathering, although they consider
the collapse model discussed above to be the more likely mechanism.
It is not clear on the timing for development of liquefaction and its progression in
relation to the stability of the slope during pre-collapse deformation and that at collapse,
particularly for the flow slides in British Columbia. At South Wales it is considered
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.49

that the slope (or part of the slope) is in a meta-stable condition resulting in the observed
pre-collapse deformation, and it is this deformation that is the trigger for liquefaction
leading to collapse of the slope.
It is considered likely that a combination of the above mechanisms has a significant
role in the development of liquefaction and flow sliding, with the dominant mechanism
differing on a case by case basis. The collapse model in association with high stress
conditions in the toe region is considered the most likely for the failures in coal waste
spoil piles in British Columbia.

3.4.3 FLOW SLIDES FROM FAILURES IN SENSITIVE CLAYS

The following discussion on flow slides in sensitive clays covers the material properties
of sensitive clays and the mechanics of failure. Two aspects of the mechanics of failure
are discussed; the initial failure (general to landslides in sensitive soft clay
environments) and factors associated with those failures that develop into “rapid”
retrogressive flow slides.

(a) Material Properties


Highly sensitive or quick clays are predominantly confined to Canada, Alaska and the
Scandinavian countries. They are fine-grained sediments of glacial origin deposited in a
marine environment. Post glacial uplift of the land mass has raised these soft marine
deposits above sea level and fresh water leaching has removed the salt ions within the
pore water. The soil structure of these leached fine-grained soils is described as
“flocculated” or open structured, and unstable, i.e., susceptible to contraction and flow
liquefaction on disturbance potentially leading to flow sliding. Torrence (1996)
comments that the sensitivity of some of these soils is enhanced by cementation, for
example the Champlain clays of eastern Canada.
Particle size distributions of sensitive clays in general and from flow slides are
shown in Figure 3.28. Fines fractions (minus 60 micron) typically range from 85 to
100% and clay fractions (minus 2 micron) from 15 to 70%.
The material and pore water properties of sensitive clays that make them susceptible
to flow sliding (Tavenas 1984; Leroueil et al 1996; Lefebvre 1996; Trak and Lacasse
1996; and Torrence 1996) are:
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.50

• The clay fraction comprises a significant content of non-clay minerals (quartz and
feldspar) derived from the grinding action of glaciers, resulting in the low activity
and generally low plasticity of sensitive clays.
• Low pore water salinity. Tavenas (1984) suggests that flow sliding occurs in
sensitive clays with salinity levels of less than 3 g/litre salts.
• A high undrained brittleness index (Figure 3.24) and very low residual undrained
strength, cur (determined from vane testing), is required for flow. Leroueil et al
(1996) suggest that for flow sliding to occur cur (vane) is less than 1 kPa or the
liquidity index, IL, is greater than 1.2. Leroueil et al (1983) proposed an empirical
relationship (Equation 3.6) relating the residual undrained strength (determined from
vane testing), cur, to the liquidity index.

cur ( vane) ( kPa) = ( I L − 0.21) −2 (3.6)

100
1 3

80
2
Percent passing (%)

60

40

20

0
0.001 0.01 0.1 1
Clay Silt Sand
Particle size (mm)
1 - Furre, Norway (Hutchinson 1961)
2 - Vibstad, Norway (Hutchinson 1965)
3 - Indicative limits of fines and clay fraction of sensitive clays (Mitchell & Markell 1974;
Bentley and Smalley 1984; Lefebvre (1996), Lefebvre et al 1991; Gregerson & Loken 1979).

Figure 3.28: Particle size distribution of sensitive clays

Leroueil et al (1996) point out that the remoulded strength determined by vane shear can
be much lower than that determined by laboratory testing and suggest the reasons for
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.51

this are strain and strain rate effects, which in laboratory tests are not sufficient to allow
for complete remoulding of the clay. An additional reason may be that in the vane shear
test the remoulded strength may be affected by load transfer of the remoulded region
onto the surrounded intact soil and therefore the test is undertaken within an isolated
remoulded region of very low confining pressure.
A summary of the plasticity, undrained strength, liquidity index and sensitivity of
the liquefied soil on the surface of rupture for each of the case studies is given in Table
3.8 (Section 3.2.7). The sensitive clays are typically of low plasticity (IP ≤ 12% in most
cases), of very soft to firm strength consistency and moderate to high sensitivity (St of
10 to 200) with liquidity index generally greater than 1.3.

(b) Initial Failure


It is evident from the literature that flow slides in sensitive soils develop from an initial
relatively small failure (Leroueil et al 1996; Lefebvre 1996) usually located within a
steep river or stream bank subject to active erosion.
The mechanism of the initial failure is considered by a number of authors (Leroueil
et al 1996; Lefebvre 1996) to occur as a drained failure of a slope triggered by a small
change in effective stress conditions, usually associated with rising groundwater.
Laboratory undrained triaxial compression test results (Aas 1981) indicate that sensitive
clays may be susceptible to static liquefaction at stress levels less than steady state (i.e.
comply with the collapse surface concept, Figure 3.7, as per contractant cohesionless
soils). Therefore, the trigger and mechanism of initial failure is potentially similar to
that for cohesionless soi1s that are contractive on shearing as idealised in Figure 3.23.
The triggering and aggravating factors for initial landsliding in sensitive clays are:
• Increasing pore pressure conditions due to snowmelt and rainfall (observed in 8 of
10 cases analysed in which the trigger was known). Tavenas and Leroueil (1981a)
present statistical data strongly correlating the frequency of landsliding in Canada,
Norway and Sweden to the months of snowmelt and climatic wet periods. For
Quebec, Tavenas (1984) indicates that the frequency of landsliding is significantly
greater in April to June, the period associated with snowmelt and high rainfall
(Figure 3.29).
• Stream erosion. Tavenas (1984) comments that landslide activity is strongly
associated with the active erosion processes of streams and rivers and along
lakeshores due to the formation of steep and high slopes of marginal stability. Locat
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.52

and Leroueil (1997) consider erosion to be an aggravating factor for initial


landsliding rather than the trigger. The forms of active erosion include vertical
erosion (down-cutting) and lateral erosion. Tavenas (1984) comments that flow
velocities in excess of 1 m/sec are sufficient to cause lateral erosion and that
transported ice blocks enhance the erosive power of streams and rivers. In the case
of the flow slide at Furre, Norway (Hutchinson 1961) an ice jam in the river resulted
in a period of very active erosion.
• Human influences such as excavation and filling.

Figure 3.29: Frequency distribution of landslides in Quebec (Tavenas 1984)

Figure 3.30: Slope height versus slope inclination for unstable slopes in eastern Canada
(Tavenas 1984)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.53

Morphological relationships of actual failures from eastern Canada are presented in


Figure 3.30. For slopes in the Scandinavian countries (Bjerrum et al 1969) field
observations indicate that at less than 20 to 21 degrees slopes are stable regardless of
height. This type of information is considered useful as a guide to the stability of
natural slopes given the complexity associated with assessment of strength properties,
hydro-geological conditions and progressive failure.

(c) Retrogressive Flow Sliding


Two types of “rapid” flow slides are observed in sensitive clays; retrogressive sliding
that occurs as a series of rotational slides or slumps of limited size (Tavenas 1984) as
shown in Figure 3.31, and translational slides (or “flake” slides) along a liquefied
surface of rupture. Cases of translational slides from the literature (Rissa, Bekkelaget,
Furre, Vibstad as examples) appear to be confined to the Scandinavian countries. This
is considered, in some cases, to be due to sediment layering (at Furre and Vibstad the
quick clays were overlain by up to 20 m of low sensitivity materials (Hutchinson 1961
and 1965)).
The criteria that predispose a slope in soft sensitive clay to retrogressive failure
(Tavenas 1984; Leroueil et al 1996; Lefebvre 1996; Trak and Lacasse 1996) are:
• There must be an initial slope failure.
• The back-scarp must be unstable in undrained loading for retrogression to occur.
From analysis of retrogressive flow slides (Mitchell and Markell 1974), Leroueil et
al (1996) suggest a stability number, Nc ( N c = γH c u where H = height of the back-
scarp, cu = undrained strength and γ = saturated unit weight), of greater than 4 for
back-scarp instability if the plasticity index, IP, is less than 10%, and Nc greater than
8 if IP is approximately 40%.
• The slide debris must have the ability to become remoulded during the failure and
flow out of the slide crater when remoulded to leave the subsequent back-scarp
exposed to maximum height. Leroueil et al (1996) comment that the energy
required for remoulding is dependent not only on cu but also on the plasticity index,
with clays of lower plasticity requiring less energy for remoulding.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.54

Figure 3.31: Retrogressive landslide in sensitive clay (Leroueil et al 1996)

3.4.4 FLOW SLIDES IN LOOSE SILTY SAND FILLS, HONG KONG

(a) Material Properties


HKIE (1998) undertook a detailed study of the properties of fills (from Hong Kong)
susceptible to liquefaction and flow sliding, and their findings are summarised as
follows:
• The materials classify as silty sands with fines contents (% passing 75 micron) of 10
to 50% and clay contents typically less than 10%. Most fills in Hong Kong are
sourced from completely decomposed granite.
• The fines are generally of medium to high plasticity. Liquid limits typically range
from 35 to 70% and plasticity index from 20 to 30%.
• The dry density of fills placed by end dumping is generally in the range 70 to 85%
of Standard Maximum Dry Density (SMDD), indicative of their loose condition.
The dumped fills tend to show layering parallel to the face slope.
• Laboratory triaxial tests indicate the loose placed fill (derived from completely
decomposed granite) is contractive at densities less than about 85 to 90% of SMDD.

Seven of the case studies analysed from Hong Kong were considered to be flow
slides in loose fill, one of which is a retained loose fill that possibly liquefied once the
initial failure had de-stabilised the retaining wall. A summary of the material properties
and slope geometry features of these landslides is given in Table 3.10 along with those
from an additional five flow slides in loose fill reported by HKIE (1998). The available
information on material type and density for these cases is in agreement with the
findings from HKIE (1998).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.55

(b) Factors Influencing Flow Sliding in Loose Silty Sand Fills


From analysis of the case studies and review of published literature it is evident that
there are several key factors that influence the likelihood of a flow slide occurring:
• Material type and material density as previously discussed.
• Rainfall. The trigger for most flow slides in loose fill slopes in Hong Kong is
rainfall. Further details on the relationship between rainfall and landsliding in Hong
Kong are given in Section 3.5.2.2.
• Fill slope angle. The fill slope angle for the flow slide case studies (Table 3.10) is
greater than or equal to 34 degrees.

(c) Failure Mechanism for Flow Sliding in Loose Silty Sand Fills
Laboratory undrained triaxial compression tests on loose silty sand fill samples of
completely decomposed granite (HKIE 1998) indicate that these soil types display
undrained strength characteristics similar to liquefaction susceptible loose sands and
silty sands (Section 3.3.2). For samples prepared at a dry density ratio of 85% of
SMDD or less and tested at low confining pressures, the peak undrained strength
defining an instability line (or collapse surface) was observed at c ′ = 3 kPa and φ ′ = 26
degrees (Figure 3.32), which is below the range of the steady state strength of 36 to 40
degrees. Therefore, the loose silty sand fills are susceptible to contraction on shearing
when in a saturated or near saturated condition at effective stress states at or above the
instability line.
HKIE (1998) indicate that the fills in the field are in a partially saturated condition,
with the degree of saturation typically ranging from 30 to 100%. In the upper several
metres of the fill slope significant changes in soil suction are measured due to climatic
influences, with near fully saturated conditions observed in the wet season due to
infiltration from rainfall. Wetting up of the soil mass and reduction in soil suction
results in a leftward shift of the effective stress state toward the instability line (Figure
3.23) within the relatively steep fill slopes, resulting in conditions under which the fill is
susceptible to liquefaction. The shallow depth of the surface of rupture (1.5 to 4 m) for
the flow slide case studies (Table 3.10) reflects the shallow depth affected by seasonal
changes in degree of saturation.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.56

Table 3.10: Features of flow slides in loose fill from Hong Kong
Slide Name Date of Material, Placement Slope Fill Slope Slide Base Angle Slide Travel
Failure Method, Dry Density Ratio Height Angle Depth of Slide Volume Distance Angle
(m) (º) (m) (º) (m3) (º)
Sau Mau Ping 25/8/76 CDG, 75 – 90% SMDD 32 34 2 to 4 34 to 40 2500 < 22
Kennedy Road 8/5/92 Silty sand (SM), CDG, placed in 15 35 2 34 500 31
1940s.
Sha Tin Heights (A) 4/6/97 Silty sand (SM) 10 34 to 49 1.5 34 150 25
Kau Wa Keng San 4/6/97 CDG, end tipped 8 36 to 40 3 to 4 20 500 23
Tseun (D)
Kau Wa Keng San 4/6/97 Silty gravelly sand (SM), loose 2.5 90 (retained) 2.5 33 to 51 60 26
Tseun (C) to very loose
Chung Shan 4/6/97 Silty sand (SM), avg 75% 14 37 3 30 450 18
Terrace SMDD
Au Ta Village 9/6/98 Silty gravelly sand (SM) 70% 7 35 2 23 170 24
SMDD
1
Kwun Tong* 1964 No details - - - - 10000 17
1
Victoria Height* 1966 No details - - - - 4000 19
1
Sau Mau Ping* 1972 No details 40 35 - - 6000 16
Princess Margaret 1989 No details 11 40 - - 200 26
1
Road *
Kowloon Wah Yan 1989 No details 12 40 - - 50 24
1
College*
*1 flow slides in loose fill from HKIE (1998) SMDD =Standard Maximum Dry Density CDG = completely decomposed granite
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.57

Further laboratory test results on loose silty sand fill samples published by HKIE (1998)
demonstrate the effect of soil suction (Figure 3.33). Increasing soil suction results in an
upward shift of the steady state line in void ratio versus mean effective stress space. At
a void ratio and mean effective stress defined by Point B a sample is susceptible to
contraction when in a saturated condition but dilative in a partially saturated condition
defined by the upper curve.

Figure 3.32: Peak shear strength from undrained triaxial compression tests of
contractive fill samples (HKIE 1998)

3.4.5 TAILINGS DAMS

The particle size distribution of tailings susceptible to liquefaction and flow from the
case studies analysed are shown in Figure 3.34. The mine tailings cover a broad range
of soil types from silty clays to clayey silts to silty sands to medium grained sands. In
most cases the tailings were derived from the crushing of ore and the finer clay and silt
sized fractions are predominantly finely ground rock particles rather than clay minerals.
Therefore, the finer particle size distributions, representative of the middle beach region,
generally classify as silts of low plasticity (plasticity index generally less than about 5 to
10%).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.58

Figure 3.33: Effect of partial saturation on steady state line in void ratio – mean
effective stress space (HKIE 1998)

To say the finer grained tailings are liquefaction susceptible is somewhat misleading
given that these materials often exist in a virtually liquid state during (and for some time
after) operation of the tailings dam, and therefore if exposed by breach will flow. The
more critical range of liquefaction susceptible soils is the silty and sandy soil types with
likely low clay contents that settle out relatively rapidly and have measurable strength
properties. The particle size distributions of the tailings are similar for the earthquake
triggered and statically triggered flow slide events.
Apart from earthquake triggered flow slides in tailings dams, the majority of the
case studies analysed (Table 3.4) were triggered by drained instability of the retaining
embankment (6 or 7 of the 9 cases). Piping was the trigger in 2 cases and over-topping
in 1 case. Once a breach in the embankment had occurred and exposed the stored
tailings, “rapid” retrogressive landsliding ensued in 8 of the 9 cases. The mechanism of
retrogression is considered to be similar to that for sensitive clays (Section 3.4.3) with
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.59

continued back-scarp instability and evacuation of liquefied tailings out of the initial
breach. Several characteristics associated with flow slides in tailings dams are:
• The tailings dam was in operation (or recent operation) at the time of failure.
• Most retaining embankments were constructed by upstream methods or a
combination of upstream and centreline methods.
• The shape of the stable back-scarp within the impoundment after failure generally
consisted of a series of near horizontal terraces with relatively low height vertical
steps separating the terrace sections. This is considered indicative of the layering
within the impoundment.
• Stratification within each layer of deposited tailings is considered to have a
significant effect on susceptibility to liquefaction. From the literature review on
liquefaction it is evident that silty sands are more susceptible to liquefaction at a
given density index and similar effective stress conditions. However, it may also
relate to water tables perched on the silty layers so the coarse tailings are not
saturated.

100

2 8
4
80
6 3
Percent passing (%)

60

9
10 9
40
5
7
20

0
0.001 0.01 0.1 1
Clay Silt Sand
Particle size (mm)
1 - Bafokeng, near discharge (Blight et al 1981) 6 - Stava, silty tailings (Chandler & Tosatti 1995)
2 - Bafokeng, from centre (Blight et al 1981) 7 - Stava, sandy tailings (Chandler & Tosatti 1995)
3 - Merriespruit, fine limit (Fourie 2000) 8 - El Cobre, earthquake triggered (Dobry & Alvarez 1967)
4 - Merriespruit, coarse limit (Fourie 2000) 9 - Mochikoshi, earthquake triggered (Marcuson et al 1979)
5 - East Texas, gypsum tailings (Kleiner 1976) 10 - Tailings slimes, low resistance to liquefaction (Ishihara 1985)

Figure 3.34: Particle size distribution of tailings from flow slide case studies.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.60

The failure at Buffalo Creek, West Virginia, USA is considered to be different to the
other failures in that it is not clear if flow sliding of the impounded slimes occurred.
The slimes may have been eroded and then transported by the large quantity of water
impounded by the dams.

3.4.6 HYDRAULIC FILL EMBANKMENT DAMS

The particle size distribution of hydraulically placed fill in embankment construction


covers a broad range of depositional soil types from silty clays to sandy gravels,
dependent on the particle size distribution of the materials used for fill and the distance
from the point of discharge. The particle size distributions for Fort Peck and Lower San
Fernando dams (Figure 3.35) partly reflect this broad range of soil types. For the case
studies analysed (Table 3.11) the liquefied zone of the hydraulic fill generally consisted
of sandy silts to silty sands to fine grained sands.
The likely mechanics of failure for the case studies is summarised as follows:
• For Lower San Fernando (LSFD) and Sheffield dams the flow slides were triggered
by earthquake. Under the stress conditions imposed by the earthquake the
hydraulically placed silty sands to sandy silts in a loose to medium dense condition
and similar foundations soils (Sheffield dam) liquefied. Castro et al (1992)
considered stratification to have had a significant effect on susceptibility to
liquefaction at LSFD.
• At Fort Peck and Calaveras dams, both flow slides during construction, abnormally
large deformations occurred prior to the flow slide event, an indication of the meta-
stable condition of the slope. At Fort Peck a significant factor in the pre failure
deformation was the basal shear deformations along pre-sheared bentonitic clay
seams in the foundation. It is suspected in both cases that these deformations were
significant in the triggering of liquefaction within the hydraulic fill and resultant
“rapid” flow slide that developed.
• Wachusett dam is somewhat different to the other cases. A flow slide developed in
the loose dumped sand fill forming the upstream shoulder of the embankment during
first filling. The very loose to loose dumped sand fill was most likely to have been
susceptible to contraction on wetting, and it is suspected that these strains associated
with wetting up of the fill were a significant factor in the triggering of the flow slide.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.61

100

2
80
Percent passing (%)

1
60
3

40

1 - Core grading limits, Fort Peck


dam (Middlebrooks 1940)
20 2 - Fort Peck dam - average grading of
shell material (Middlebrooks 1940)
3 - Zone of liquefaction, Lower San
Fernando dam (Castro et al 1992)
0
0.001 0.01 0.1 1 10
Clay Silt Sand Gravel
Particle size (mm)
Figure 3.35: Particle size distributions of hydraulic fills from flow slide case studies

Table 3.11: Summary of properties of liquefaction susceptible materials from flow


slides in hydraulic fill embankments
Name Material Properties Embankment Comments on Trigger and Slide
(Liquefiable Zone) Height Slope Location
(m) (º)
Sheffield Dam Silty sands to sandy silts in a 7.6 22 Slide in downstream shoulder triggered
loose to medium dense by earthquake. Liquefaction of either the
condition (DR = 35 to 50%). lower portion of the embankment or the
foundation.
Lower San Silty sands to sandy silts in a 35 21.8 Slide in upstream shoulder triggered by
Fernando medium dense condition (DR earthquake some 46 years after
Dam = 50 to 55%), highly construction.
stratified.
Calaveras Highly variable depositional 56 18 Slide in upstream shoulder during
Dam PSD from sandy gravels to construction. Hydraulic fill from
silty sands to silty clays in colluvial soils and soft sandstones.
the pool area. Likely that siltier and sandier soil types
liquefied.
Fort Peck Fine to medium sand, 58 12.6 Slide in upstream shoulder during
Dam stratified construction. The likely trigger of
liquefaction was deformation along
bentonitic seams at depth in the shale
foundation.
Wachusett Fine sand to silty sand in a 24.4 22 Slide in upstream shoulder during first
Dam * 1 loose to very loose filling. Liquefaction of loose dumped
condition. Placed by fine sands, most likely due to strains on
dumping. saturation.
*1 Constructed of loose dumped sand
DR = relative density PSD = particle size distribution
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.62

3.4.7 SUBMARINE SLOPES AND SLOPES IN SUB-AQUEOUS FILLS

The particle size distributions from flow slide events in coastal sub-aqueous slopes and
sub-aqueous fill slopes (Figure 3.36) indicate that fine to medium grained sands with
low fines contents (less than 10%) are susceptible to static liquefaction (as opposed to
earthquake or other dynamic trigger) and flow sliding. The spectrum of particle size
distribution in these environments is broader than shown in Figure 3.36 most likely
ranging from soils with relatively high silt contents (Karlsrud and Edgars 1982) to
sandy gravels (Howe Sound, Terzaghi (1956)).
For flow slides covering the broader spectrum of submarine slopes and including
earthquake triggered flow slides Edgars and Karlsrud (1983) indicate flow slides, whilst
predominantly in silts and fine sands, have been observed in more clayey and more
gravely soils.

100
2 3

80 2
2 2
Percent passing (%)

60
4
1

40

20

0
0.01 0.1 1 10
Silt Sand Gravel

Particle size (mm)


1 - Flow slides in coastal sub-aqueous slopes, Zeeland (Koppejan et al 1948)
2 - Flow slides in coastal sub-aqueous slopes (Kramer 1988)
3 - Sub-aqueous fill berm, Nerlerk sand (Sladen & Hewitt 1989)
4 - Sub-aqueous fill slope, Willamette River (Cornforth et al 1974)

Figure 3.36: Particle size distributions of soil types from submarine slopes and sub-
aqueous fills susceptible to liquefaction and flow sliding.

Relative density is also a significant material property of sub-aqueously deposited soils


that are susceptible to liquefaction. As discussed in Section 3.3.3 the relative density
defining the liquefaction boundary is dependent on material type and effective stress
conditions. For the case studies in sub-aqueous fill slopes (Table 3.6) flow sliding
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.63

occurred in very loose, medium to coarse sands (Willamette River) and very loose to
loose, fine to medium grained sands (Nerlerk Berm), both with low silty fines contents.
For flow slides in coastal sub-aqueous deposits a common element of the failures is
that they occur in geologically recent deposits generally placed under rapid
sedimentation conditions. In Norway post glacial deposits near river mouths are
susceptible (Karlsrud and Edgars 1982) and in the Zeeland province of The Netherlands
tidal channel fills and rapidly deposited fluvial channel fills of sand are identified as the
most susceptible to flow sliding (Silvas and de Groot 1995). Silvas and de Groot
(1995), as a broad guide to assessment of susceptibility to liquefaction (for Zeeland
province), consider loose sands from CPT testing or sands with a relative density of less
than 30% to be susceptible to liquefaction and flow sliding, and sands with a relative
density of greater than 60% as not susceptible.
The mechanism associated with initial failure and retrogression is similar to that for
sensitive clays. Silvas and de Groot (1995) consider that three conditions are necessary
to trigger flow sliding:
• The soil must be susceptible to liquefaction;
• The slope must be relatively steep and relatively high for the initial failure to be
triggered. From empirical data (for Zeeland province) they suggest a steep section
within the overall slope greater than 5 m in height and steeper than 18.4 degrees, but
consider this to be conservative due to the simplistic qualitative evaluation; and
• An initiation or triggering mechanism must be present. Identified triggers have
included changes in water pressure due to waves, increased outflow of groundwater
during extremely low spring tides, dredging activities, erosion at the toe (causing
localised over-steepening) or vibration.

Edgars and Karlsrud (1983) considered either earthquake and/or rapid sedimentation
to be the trigger in most cases they analysed.
In effect, therefore, the slope conditions provide the effective stress conditions under
which the soil is susceptible to liquefaction (i.e. at or above the instability line, Figure
3.23). The static trigger to flow sliding is then a change in effective stress conditions
(Section 3.4.1).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.64

3.5 M ECHANICS OF F AILURE IN DILATIVE SOILS LEADING TO “RAPID”


SLIDING
This section discusses the mechanics of failure for failures of slopes in dilative soils that
developed into “rapid” landslides. Within each landslide sub-group the properties of the
slide material and factors influencing the triggering of and development of slope failure
are drawn from the literature and case study data.

3.5.1 SLOPE TYPES OF FAILURES IN DILATIVE SOILS LEADING TO “RAPID ” SLIDING

“Rapid” landslides from slope failures in dilative soils occur in steep cuts in residual
soils, completely weathered rock and colluvium, in natural slopes of mostly colluvium,
and steep compacted fill slopes. Retained soil slopes are another class of slope within
which failures develop into “rapid” landslides, but they have not been considered in
detail here.
From the case study data, the slides in dilative soils are mostly either translational or
have significant translational components. For slides in colluvium and compacted fills
the surface of rupture is generally through the soil mass and is, in a number of cases,
coincident with soil interfaces either between the colluvium and underlying weathered
rock or between differing soil units in the colluvium. For slides in residual soils and
weathered rocks the surface of rupture (or part thereof) may preferentially be located
along defects within the slide mass that are adversely oriented to the face slope and the
strength along which is significantly lower than that of the soil mass. In the following
discussion defect controlled slides are considered separately to slides through the soil
mass in dilative soils. Only those slope failures where there is evidence of strong
defect-control are grouped under the defect-controlled class of dilative slide; all others
are included under the slides through soil mass class (including the dilative fill slope
failures).
To be consistent with the landslide classification system (Figure 1.1) and avoid
confusion with terminology, the term “slide of debris” is used to describe the initial
landslide in preference to the terms “debris slide” and “debris flow” used by some of the
referenced authors, which are more appropriate to the post failure travel of the slide
mass.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.65

3.5.2 SLIDES THROUGH SOIL MASS – DILATIVE FAILURES

The literature on slides of debris presented in this chapter is sourced from published
research from Hong Kong as well as that from the international community. The strong
emphasis on the research from Hong Kong is because of the availability of detailed
reports on individual landslides and studies into rainfall and topography associated with
slides of debris.
Also included within this section are slides through the soil mass, where the soil is
initially dilative on shearing, for failures in cut slopes in residual soils and deeply
weathered rocks, and compacted fills where the post-failure travel was “rapid”. The
case study information on these failure types is from Hong Kong.

3.5.2.1 Mechanics of Development of Dilative Slides Through the Soil Mass

Debris slides and debris flows in natural slopes initiate from either landslides (slides of
debris) or as slurry that evolves into a debris flow through channel erosion (Johnson and
Rodine 1984). Only those “rapid” debris slides or debris flows initiated from a
landslide are considered here.
The typical geometric features of slides of debris in natural slopes (Hutchinson
1988; Corominas 1996b) are shallow failure depth, low depth to length ratios (often <
0.05) and high length to breadth ratios (typically 5 to 10 or greater).
An important aspect of landslides from slides of debris (and other slide types) is the
distinction between the initial development of the failure itself and the transformation of
the slide mass into a “rapid” debris flow or debris slide. Initiation of the failure occurs
when a critical equilibrium condition is reached within the slope along a defined surface
of rupture. Triggering of an unstable slope condition is dependent on material strength,
slope geometry and the hydro-geological conditions within the slope. Often it is
changes in hydro-geological conditions due to rainfall or snowmelt and their effect on
material strength properties due to changes in soil suction that trigger the initial slide.
Most authors are in general agreement with the fundamentals of this process (Johnson
and Rodine 1984; Brand et al 1984; Corominas 1996b; Iverson et al 1997; Fannin and
Rollerson 1992; amongst others).
The subsequent models to explain the mechanics associated with transformation of
the slide mass from a “slide of debris” into a “rapid” debris flow or debris slide is
largely based on observation of actual flow slides or large-scale experimental work
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.66

within a soil mechanics framework. Broadly speaking there are two main models; one
is associated with contractile behaviour on shearing and subsequent liquefaction related
flow sliding, and the other associated with dilative behaviour of the soil mass. In this
section the dilative response is of interest, but the observations of liquefaction flow
slides in natural slopes of colluvium (i.e. slides of debris) will be briefly discussed.
Large-scale experimental studies reported by Iverson et al (1997) indicate that in
contractile soils the progression from initial failure to “rapid” flow occurs very abruptly
and without warning. In the field a similar observation of very abrupt transformation of
a slide of debris (in soils that are contractive on shearing) to a flow condition is
observed, particularly for very shallow slides in the surficial colluvium (Ellen et al
1988). From the previous discussion on flow slides in Section 3.4.1 (within which these
slides in contractile colluvium would be classified), the progression of the liquefied
zone along the surface of rupture is considered to occur very rapidly once initiated and
therefore transformation to a “rapid” flow is very abrupt.
For slides of debris in soils that are initially dilative on shearing, the mechanics
associated with transformation of the slide mass into a “rapid” debris flow or debris
slide is considered by several authors (Johnson and Rodine 1984; Fleming et al 1989;
Corominas et al 1996) to involve dilation and remoulding of the slide mass on shearing.
Johnson and Rodine (1984) idealise the transformation process once a failure condition
has been reached as:
• Progressive failure starts from the base of the slide area where deformation, dilation,
tensile cracking and incorporation of water occur, leading to sliding of the mass.
• Landsliding and remoulding progress, whilst the debris continues to soften due to
dilation, water ingress and remoulding.
• The slide mass loses coherency and begins to slide (or flow).

The observations of development of slides of debris (Johnson and Rodine 1984)


indicate that the transformation of the slide of debris into a “rapid” debris slide or debris
flow occurs over distances of 1 to 3 m (sometimes in a matter of seconds), with a rapid
change in velocity initially from very slow to “rapid”. They comment that the strain
softening during transformation is critical to the development of “rapid” movement of
the slide mass.
Corominas et al (1996), whilst generally agreeing with the initiating mechanism
described by Johnson and Rodine (1984), comment that it is uncertain. They believe
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.67

liquefaction may develop within the base of the slide mass (post the initial failure),
giving rise to the rapid increase in velocity observed.
Iverson et al (1997) observed, from field experimentation, that for failure in a
dilative soil the slide mass did not take on the appearance of a liquefied soil mass until it
had evacuated the source area and was descending the 50 degree run-out ramp.
Interestingly, the soil moisture content at failure and during flow was similar between
the slides in the contractile and dilatant soils in their experiments (both soils were of the
same type). Iverson et al (1997) concluded that for the dilative soils, at some point in
the travel of the landslide mass beyond the source area excess pore pressures are
developed due to conversion of energy within the slide mass leading to transformation
into “rapid” debris flow. Another important observation by Iverson et al (1997) was
that additional ingress of water was not required for remoulding.
Fleming et al (1989) investigated the Salmon Creek slide of debris (Marin County,
California) and concluded that the initial slide, predominantly within a dilative clayey
silty sand colluvium, transformed into a debris flow. They observed a basal surface of
rupture of about 10 cm thickness, located within the colluvium directly above the
weathered sandstone, that comprised a series of softened shear surfaces. Density
measurements within the basal shear zone indicated that the material had dilated about
5% in volume compared with similar material outside the failure area, and that the mass
above the basal shear zone had also dilated. Fleming et al (1989) considered that the
slide of debris within the initially dilative colluvium had transformed into a contractive
material (Figure 3.37) through dilation of the slide mass leading up to the development
of the eventual debris flow.
The mechanism proposed by Fleming et al (1989) is considered a reasonable
mechanism for slides in dilative soils on natural slopes.
For slides through the soil mass in cut slopes, the loss of strength from peak as the
slide displaces is sufficient to explain the acceleration of the slide mass. If the cut slope
is steep, then the post failure movement of the slide mass is generally “rapid”.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.68

Figure 3.37: Hypothetical model for transformation of dilative materials from rigid to
fluid behaviour, path Di to Df (Fleming et al 1989)

3.5.2.2 Correlation of Rainfall to Triggering of Slides of Debris

Rainfall is considered by most authors (Johnson and Rodine 1984; Hutchinson 1988;
Ellen et al 1988; Brand 1989; Kim et al 1991; Wong et al 1996; Corominas 1996b;
Iverson et al 1997; amongst others) to be the trigger for initiation of most slides of
debris, but other sources include snowmelt and springs.
In areas where a high incidence of rainfall triggered landsliding occur with the
potential to cause loss of life and damage to property, studies are generally undertaken
in an attempt to correlate rainfall and landsliding for the purpose of providing warning
systems. A common finding by authors researching this relationship (Caine 1980;
Wieczorek 1987; Cannon and Ellen 1988; Kim et al 1991; Church and Miles 1987;
amongst others) is that both antecedent rainfall and short term rainfall intensity are
important factors. In the San Francisco Bay region for example, Wieczorek (1987)
found that sufficient antecedent rainfall was required before landslides that developed
into debris flows could be triggered, and Cannon and Ellen (1988) concluded that once
antecedent rainfall conditions were sufficient to near saturate the surficial soil profile
intense rainfall will lead to debris flow activity. Corominas et al (1996) commented that
the general form of the failure threshold concept proposed by Caine (1980), Equation
3.7, was a useful correlation that incorporated the need to consider soil saturation.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.69

I = 14.82 ⋅ D −0.39 (3.7)

where I is the rainfall intensity in mm/hour and D is the rainfall intensity in hours.
Johnson and Rodine (1984) considered it ridiculous to relate landsliding or debris
flow initiation to a single parameter such as slope geometry or rainfall intensity.
In Hong Kong, the trigger for most landslides (including slides of debris, cut slope
failures, flow slides and defect–controlled slides) has been identified as rainfall (Brand
1989, 1995). Of the slides of debris in natural terrain (Wong et al 1996; Evans et al
1997; Franks 1996) and the failures in cut slopes on Lantau Island (Wong and Ho 1996)
the trigger was identified as intense rainfall. Brand concluded that:
• Localised short duration rainfalls of high intensity induce the large majority of
landslides, and that these landslides take place at about the same time as peak hourly
rainfall.
• Antecedent rainfall is not a major influence on landslide occurrence, except in some
cases of minor landslide events.
• A rainfall intensity of about 70 mm/hour is about the threshold value at which
landsliding occurs (Figure 3.38), with the number of landslides and severity of
consequences increasing as the peak hourly rate of rainfall intensity increases above
the threshold of 70 mm/hour. Figure 3.39 shows the relationship between severity
of landslides and rainfall intensity. Note that these landslides are predominantly
from constructed fills, cuts and retained slopes, not natural slopes; and the casualty
rate is historic and unlikely to be the current situation as Hong Kong has, and is
continuing to, expend large sums in slope remedial works.

The landslide warning issued by the Hong Kong Government (obtained from the
Hong Kong Government web site) is when the 1-hour intensity exceeds 70 mm/hour
and the 24-hour intensity exceeds 175 mm. Both these warning criteria are closely
coincident with the change from minor to severe landslide events (Figure 3.39).
Brand (1989) considers that the high correlation of landsliding and short duration
rainfall is related to the relatively high permeability of the colluvial and saprolitic soils
mantling the steeper slopes in Hong Kong and the dominance of short-term rainfall
intensity is related to the rapid rise in piezometric pressures (Brand 1995). He considers
the increases in piezometric pressures observed are largely the result of rapid rises in
transient perched water and groundwater fed mainly by natural erosion pipes and
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.70

tunnels caused by sub-surface erosion observed in the soil profile at many landslides.
Franks (1996) reported that erosion pipes were observed in 65% of the failure scarps
inspected with most of the soil piping observed in residual soils and colluvium, and
Irfan (1997) observed erosion pipes in a number of saprolitic soil slopes. Brand (1989)
suspects that similar correlations to rainfall intensity would be appropriate for soils with
relatively high permeability in tropical areas. For soils of lower permeability he expects
antecedent rainfall to become increasingly important.

Figure 3.38: Relationship between peak hourly rainfall, landsliding and severity of
consequence for Hong Kong (Brand et al 1984)

Figure 3.39: Relationship between severity of landsliding and rainfall intensity for Hong
Kong (Brand 1985b)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.71

It should be noted that Finlay et al (1997), using detailed rainfall and landslide data for
Hong Kong Island that allowed better correlation of rainfall and sliding than the
Territory wide data Brand (1989, 1995), demonstrated that antecedent rainfall did have
some influence. This is also evident for the case studies analysed from Hong Kong
where the depth to sliding is greater than about 5 to 7 metres and the slide volume is
greater than about 1500 to 5000 m3.
Of the forty-three landslides analysed from Hong Kong (refer Section 1 of Appendix
B) the trigger in forty-two cases was rainfall, with one case (Kwun Lung Lau on 23 July
1994) triggered by a leaking service pipe. From analysis of these case studies the
following tentative conclusions are drawn:
• For the most part, the significant rainfall triggering landsliding was rainfall
intensities over periods ranging from less than 1 hour up to 4 hours. This is
particularly significant for landslides of shallow (= 3 m depth) to medium depth (>
3 m to = 7 m depth).
• Longer term antecedent rainfall in conjunction with heavy rainfall at the time of
sliding appears to be significant for some of the deeper and larger volume
compound and translational defect controlled slides (Po Shan Road, Fei Tsui Road,
Ten Thousand Buddhas Monastery, Ma On Shan Road and Shum Wan). The 15 to
31 day rainfall events were significant for these landslides with reported annual
probabilities of 1 in 40 years to 1 in 1000 years.
• Delay between heavy rainfall and landsliding was noted in three cases (Castle Peak
Road - Milestone 14.5, Lido Beach and Allway Gardens). In all three cases the
GEO reports considered that rising groundwater levels triggered landsliding (as
opposed to development of transient perched water in most cases).

Vegetation affects the frequency of debris slides and debris flows in several ways.
Positive effects of vegetation in reducing the frequency of landsliding include;
reduction in the amount of precipitation reaching the ground surface, reduction in soil
moisture content through transpiration, and the root system may strengthen the upper
soil profile. However, vegetal cover does increase infiltration compared with barren
soil because it retards surface water flow and the root system makes the soil more
pervious. Hutchinson (1988), Corominas et al (1996) and Johnson and Rodine (1984)
consider that the destruction of fire to be significant in increasing the probability of
debris slides and debris flows. From a study in North Lantau, Hong Kong, Franks
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.72

(1996) commented that sparsely vegetated slopes appeared to be the most susceptible to
landsliding. Evans et al (1997) found it difficult to draw any conclusions with respect
to slope vegetation.
It is possible that the implicit assumption drawn from the above discussion is that
debris slides and debris flows are only associated with intense and/or antecedent
rainfall. This is not correct. Rather it is rainfall that initiates sliding, then the soil and
slope characteristics that determine whether the movement of failed slide mass is
“rapid” or will transform into a “rapid” debris slide or debris flow.

3.5.2.3 Material Properties

Indicative particle size distributions and material properties of the source area of slides
of debris in natural slopes that developed into “rapid” debris slides and debris flows
from published literature are presented in Figure 3.40 and Table 3.12. Several of the
particle size distributions have been estimated from material descriptions given by the
referenced author/s. The source area material types in natural slopes susceptible to
development of “rapid” sliding cover a wide range from high plasticity silts to low to
medium plasticity clays and clayey sands to coarse grained granular soils
(predominantly cobble to boulder size) with low fines contents. The broad range is
considered to reflect the significant influence of slope geometry.
Two modes of post failure deformation and slide mass behaviour are evident for the
slides in dilative soils:
• Slides of debris that remain relatively intact during travel (i.e. debris slides). For the
slides in the San Francisco Bay region Ellen et al (1988) comment that the higher
clay content soil types (high plasticity silts and low to medium plasticity sandy clays
and clayey sands) generally remain relatively intact during travel and reach
maximum slide velocities of metres per second (rapid to low end of very rapid
category). The slides at La Honda (Figure 3.40 and Table 3.12) are an example.
• Slides of debris that undergo significant break up and remoulding, transforming into
debris flows. Material types typically range from clayey silty sands to coarse
grained granular soils (predominantly cobble to boulder size) with low fines
contents. Travel velocities are much higher than for the debris slides in finer
grained and more plastic soils, and typically in the range very rapid to extremely
rapid. For the slides in the San Francisco Bay region Ellen et al (1988) comment
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.73

that the higher velocity slides are observed in soil types with an upper clay content
of 25%. The published data (Ellen et al 1988) also indicates that these soils have
low plasticity index (0 to approximately 10 to 15%).

100

2 3 1
80
1
1
Percent passing (%)

2 1 1
4
60

5
40

20
1 6
7

0
0.001 0.01 0.1 1 10 100
Clay Silt Sand Gravel
Particle size (mm)
1 - San Francisco Bay region (Ellen et al 1988; Johnson & Rodine 1984; Fleming et al 1989)
2 - La Honda, San Mateo County, San Francisco (Ellen et al 1988)
3 - Alameda County, San Francisco (Ellen et al 1988)
4 - Heath Canyon, Los Angeles (Johnson and Rodine 1984)
5 - Colluvium, Hong Kong (case studies), grading estimated from description
6 - Colluvium, Japan (Ikeya1989), grading estimated from description
7 - Vancouver & Queen Charlotte Islands, Canada (Fannin et al 1996), grading estimated from
description

Figure 3.40: Source area particle size distributions of dilative slides of debris in natural
slopes.

For slides of debris in natural terrain from Hong Kong the colluvial soil types (in the
source area) ranged from silty sands to predominantly cobble to boulder size materials
with clay to gravel contents as low as 15 to 25%. Failures in natural slopes (from Hong
Kong) have also been reported to incorporate residual and saprolitic soils of granitic and
volcanic origin classifying as silty sands to clayey silty sands to sandy silts with varying
gravel to boulder size content.
Rodine (1974) considered that the reason why the transformation into debris flows
were uncommon in clay-rich materials was because of the difficulty of introducing
enough water into soils of relatively low permeability to reduce the strength sufficiently
to form a debris flow. Johnson and Rodine (1984) add that debris flows can develop in
clay-rich soils when the soil has a high degree of saturation (and low strength) and
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.74

requires only a small increase in moisture content for instability and sufficient softening
to allow a slide or flow to develop.
Johnson and Rodine (1984) comment that slides of debris that transform into “rapid”
debris slides and debris flows are significantly more prevalent in cohesionless or low
cohesion soils as they are more permeable and require much smaller changes in soil
moisture for strength reduction and remoulding. They add that in general, debris flows
(as opposed to debris slides) comprise only minor amount of clay with the larger clasts
typically carried (or suspended) on or within the debris flow. Corominas et al (1996)
describe the debris flow material type as typically well graded with clay contents of less
than 5%.
Material permeability is also a significant material property affecting the
development of slope instability and the transformation from landslide into debris
flows. Numerous authors (Johnson and Rodine 1984; Brand 1989; Brand 1995; Franks
1996; Iverson et al 1997; amongst others) comment that soil permeability affects the
rate and depth to which water infiltrates the ground, and that defects within the soil
mass (erosion pipes, root holes, etc.) and differential material permeability enhance
slope failure by development of locally elevated pore pressure conditions or transient
perched water.
The range of particle size distribution for slides of debris that develop into “rapid”
debris slides and debris flows is very broad. The grading limits in Figure 3.40 and
material properties in Table 3.12 and the above discussion are considered useful as a
guide in conjunction with assessment of slope geometry and other factors affecting the
transformation from slides of debris in “rapid” debris slides and debris flows.

3.5.2.4 Slope Geometry


A summary of the source area slope angle (a 1 in Figure 1.2) of dilative slides that
developed into “rapid” debris slides and debris flows from this study and published
sources is given in Table 3.13. Table 3.14 summaries the source area and initial down-
slope slope geometry for the case studies (from Hong Kong) considered in this study.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.75

Table 3.12: Summary of material types of dilative slides of debris in natural slopes that developed in “rapid” debris slides and debris flows.
Location/Region Material Properties Slide Description References
Classification Properties
Marin County, San SM/ML - clayey silty sands Low plasticity, WL = 18 to 24%, Slides of debris that developed in “rapid” Fleming et al (1989)
Francisco to sandy silts, colluvium IP = 0 to 4% debris flows. Very to extremely rapid. Ellen et al (1988)
La Honda, San Mateo MH High plasticity Travelled as relatively intact debris slides. Ellen et al (1988)
County, San Francisco “Rapid” velocities in metres/minute.
Pacifica, San Mateo SC/CL – sandy clays and Low plasticity, WL = 18 to 42%, Slides of debris that developed into debris Ellen et al (1988)
County, San Francisco silts, and clayey sands IP = 3 to 20% flows. Rapid to extremely rapid. Howard et al (1988)
Heath Canyon, Los Sandy gravel to gravelly Low fines content (< 10%) Slide of debris that developed into debris Johnson and Rodine
Angeles sand flow (1984)
Vancouver and Queen Colluvium. Gravelly sands Some to little silt and trace clay Slides of debris that developed into debris Fannin et al (1996)
Charlotte Islands, to sandy gravels. flows.
Canada
Queen Charlotte Islands, Colluvium. Silty sands to Some to little silt and trace clay Slides of debris that developed into debris Fannin and Rollerson
Canada sandy silts. flows. (1992)
Taeback Mountains, Residual soils (and colluvium?) derived from gneiss, schist Developed into debris flows. Kim et al (1991)
Korea and granitic rock types.
Eastern Alps, Austria Moraine. From boulder to clay size, loose. Developed into debris flows. Very to Aulitzky (1989)
extremely rapid.
Japan Sandy gravels to gravelly sands with boulders and low fines Developed into very rapid to extremely Ikeya (1989)
content (< 10% passing 0.1 mm) rapid debris flows.
Campanian Apennines, Colluvium derived from pyroclastic soils. Sandy gravels to Slides of debris that developed into Guadagno (1991)
Italy clayey sandy silts. Low plasticity (WL = 25 to 34%, IP close confined debris flo ws.
to 0%).
WL = liquid limit IP = plasticity index
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.76

For slides of debris in natural slopes Hutchinson (1988), Corominas et al (1996) and
Corominas (1996b) consider that slope angles in which slides of debris are transformed
into “rapid” debris slides and debris flows typically range from 25 to 45 degrees,
although they can take place on slope angles as low as 18 degrees and as steep as 50
degrees. The findings from other studies (Table 3.13) concur with these findings.
Evans et al (1997) report the findings from more than 26,000 debris slides and
debris flows in natural terrain from Hong Kong. The study included some 18,000 relict
landslides (pre 1945) and almost 9000 recent landslides (1945 to 1994). The
distribution of head slope angle (or source area slope angle, a 1) for both the relict and
recent landslides from this study is given in Figure 3.41. As an indication of the lower
limit of slope angles in which debris slides occur in Hong Kong, Irfan and Woods
(1998) report slow moving landslides of colluvium (sliding on weathered rock interface)
on slopes of 15 to 18 degrees and in weathered saprolite soils of 21 degrees.
The locality of the initial slide is reported by the referenced authors (Table 3.13) to
be dominantly on either open slopes or the head region of the watershed. Ellen et al
(1988) commented, for the initial slides in the San Francisco Bay region during the
intense rainstorm of January 1982, that the initial slide occurred in areas conducive to
concentration of water such as concavities, breaks in the slope or geological contacts,
but also occurred in areas without these features (i.e. open slopes).
From Table 3.13 it can be concluded that development of “rapid” debris slides and
debris flows derived from slides of debris in natural slopes predominantly occur where
the source and immediate down-slope angle is greater than about 25 degrees, but can
occur on slopes down to about 18 to 20 degrees. The broad regional spread would
suggest that these slope angles are typical in most regions of high landslide intensity
around the world.
For the cut slope case studies from Hong Kong (Table 3.14) “rapid” slides from
failures in dilative soil types (and not defect controlled) occur in slopes greater than
about 34 degrees. This is likely to be regionally specific considering the limited
geological environment of the case studies, but is considered applicable to cut slopes in
similar soil and weathered rock types (silty sands to sandy silts with low clay contents
and decomposed granitic and volcanic rock types). Studies on cut slopes in high
plasticity clays (James (1970) and Chapter 5) indicates “rapid” sliding occurred for
failures in cut slopes steeper than about 27 to 30 degrees, with some dependency on the
location of the slide within the slope.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.77

Table 3.13: Source area slope angle ( α 1 ) of dilative slides in natural slopes that
developed into “rapid” debris flows and debris slides
Reference / Location Locality and/or Spread Source area slope
angle, a 1 (º)
This study (Hong Kong) – natural slopes In colluvium 31 to 44
In saprolite 33 to 52
Evans et al (1997) – Natural slopes in Hong Total range 11 to >45
Kong 95% of slides >27
Most landslides (75%) 30 to 45
Irfan and Woods (1998), natural slopes in Approximate minimum 15 to 21
Hong Kong
Corominas (1996b), slides of debris Total range 18 to 50
Majority 25 to 45
Hutchinson (1988) –slides of debris Typical range 25 to 45
Johnson and Rodine (1984), natural slopes in Typical range 25 to 40
USA southern California >39
Virginia, USA >17
Hawaii 42 to 48
western USA >30
Ellen and Fleming (1987), Ellen et al (1988) Typical range 25 to 40
– natural slopes in San Francisco Bay Region Approximate minimum 18 to 20 (possibly as low
as 14)
Fannin and Rollerson (1992) – Vancouver Total range > 22 to 43
and Queen Charlotte Islands Majority of slides 28 to 40
Kim et al (1991) - Taeback Mountains, Typical range 30 to 35 (some on
Korea shallower slopes)
Guadagno (1991) - Campanian Apennines, Typical > 25
Italy Approximate minimum as low as 18
1
Rapp and Nyberg (1981), Rapp (1986) * - 25 to 40
1
Zimmermann (1990) * - 25 to 40
Lewin and Warburton (1994) * 1 - 15 to 39
1
Campbell (1975) * - 22 to 45
1
Johnson and Rahn (1970)* - 25 to 40
*1 Sourced from Corominas et al (1996)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.78

Table 3.14: Failures through the soil mass in dilative soils that developed into “rapid”
landslides, from Hong Kong
Slide Name Slope Source Initiating Slide Geometrical *1 Vol. TDA,
Type Slide Description Slopes (degrees) (m )3
a *1
Material α1 α base α2 (º)
Keng Hau Rd-B Cut Residual (G) Non-circular, int 53 29 40 210 26
Cheung Shan Cut Colluvium Slide of debris, 51 36.5 51 35 36
Estate translational, int
Lai Cho Road Cut Colluvium & Non-circular, int 34 34 34 33 32
saprolite (G)
St. Joseph’s Cut Saprolite (G) Non-circular 65 65 0 25 44
Yue Sun Garden-B Cut Saprolite (V) Compound, trans 52 50 2 1100 29
Castle Peak Rd (M Cut Saprolite (G) Rotational 46 - 0 300 28
14.5)
Yue Sun Garden-A Cut Saprolite (V) Non-circular 56 - 56 200 36
Bayview Gardens Cut Saprolite (V) Rotational 66 - 0 70 47
Ha Wo Che Shatin Cut Saprolite (G) Non-circular 55 - 0 60 34
Keng Hau Rd-A Cut Residual & Compound 64 31 26 150 19
saprolite (G)
Chung Shan Cut Saprolite (G) Non-circular 52 35 0 85 41
Terrace-B
Hong Tseun Road Cut Saprolite (V) Non-circular 36 - 43 250 25
Tao Fung Shan Nat Residual (G) Translational, int 50 30 40 900 30
Cemetery-A
GP Christian Nat Residual (G) Compound, trans, int 33 0 32.5 20 28
Centre
Lui Pok School Nat Colluvium Slide of debris, 31 31 29 450 19
compound, trans, int
Tsing Shan Nat Colluvium Slide of debris, 40 40 33 13000 23
compound, int
Lantau Island – C1 Nat Colluvium Slide of debris, 33 30 23 1000 21
compound, trans, int
Wonderland Villas Nat Residual (G) Compound, trans, int 40 37 34 80 29
Sha Tin Height – Nat Colluvium Slide of debris, 37 40 13 170 18
Slide B translational, int
Ma On Shan Rd Nat Colluvium & Compound 34 32 to 45 7 3000 22
saprolite (G)
Outward Bound Nat Colluvium & Non-circular 44 - 33.5 900 27
School saprolite (V)
Ka Tin Court, Nat Saprolite (G) Compound, trans 52 36 37.5 150 29
Shatin
Lido Beach Nat Residual & Non-circular 52 - 17 750 21
saprolite (G)
Tuen Mun Road Fill Silty sand Compound, trans 34 29 0 200 27
1
* refer Figure 1.2 for the definitions of slope geometry terms, a1 = source slope angle, abase = rupture
surface inclination angle, a2 = initial down-slope angle and TDA (or a) = travel distance angle.
Nat = natural (G) = granitic origin (V) = volcanic origin int = sliding on interface
trans = significant translational component Vol. = slide volume
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.79

35

30

25
Landslides (percent)

20

15

10

0
< 11 11 - 15 15 - 18 18 - 22 22 - 27 27 - 30 30 - 34 34 - 39 39 - 45 > 45
Head Slope Angle (degrees)

Figure 3.41: Distribution of head slope angle (source area slope angle) for slides of
debris in natural terrain that developed into “rapid” landslides, Hong Kong (Evans et al
1997)

3.5.3 DEFECT CONTROLLED LANDSLIDES

The nine case studies of defect-controlled landslides analysed are from Hong Kong.
Glastonbury and Fell (2000) in their analysis of large (mostly greater than 1 million
3
m ) “rapid” failures in rock slopes, found that for compound rupture surfaces, brittle
rock mass failure is required to cause “rapid” movement. For toe buttress compound
slides rupture of the rock mass at the toe is required, for bi-planar and curved compound
rupture surfaces brittle internal deformation is required. For translational planar
surfaces of rupture, brittleness associated with either the basal rupture surface or lateral
rupture surfaces is required for “rapid” sliding. They compared the rupture surface
inclination angle to the basic friction angle of the controlling defect for compound and
translational failures, and found that:
• For compound rock slides that develop into “rapid” landslides the rupture surface
inclination angle is typically 10 to 30 degrees greater than the basic friction angle,
indicating the significance of the rock mass restraint on “rapid” sliding.
• For translational rock slides that develop into “rapid” landslides the difference
between the rupture surface inclinational angle and the basic friction angle is
significantly less, generally in the order of 0 to 10 degrees, suggesting therefore that
dilation angles are of the order 0 to 10 degrees.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.80

Whilst the defect-controlled failures that developed into “rapid” landslides from
Hong Kong are of significantly smaller volume and the intact strength of the saprolitic
soils significantly less than for the rock slides analysed by Glastonbury and Fell (2000),
it is considered that similar principals are applicable for the development of “rapid”
post-failure movement for this slide group. Table 3.15 summarises the features of
defect controlled compound slides and Table 3.16 defect controlled translational slides
case studies that developed into “rapid” landslides post failure. Estimates of basic
friction angle and residual friction angle on the controlling defect are based on
laboratory shear strength test results where available from the site or else approximated
from published strength test results (Irfan and Woods (1998); Irfan (1997)) on
discontinuities within similar rock types with similar infilling. No estimate of the
asperity roughness (dilation angle) has been made.
Saprolitic soils derived from weathering of volcanics and granite in Hong Kong
typically classify as silty sands with varying gravel to boulder content depending on the
degree of decomposition (weathering) and generally have low clay contents (typically
less than about 10%). The thickness of these weathered rocks is variable; Irfan (1997)
reports thicknesses of up to 60 m in granitic rocks and typically less than 15 m in
volcanic rocks. Weathering is concentrated on structural and lithological discontinuities
in the original rock mass (Irfan and Woods (1998)) leaving relict discontinuities in the
saprolitic soil mass. These relict discontinuities often contain a thin veneer of silty clay
to clay with kaolinite often the dominant mineral. The surfaces can be rough, polished
or in some cases slickensided. These relict discontinuities are typically of significantly
lower strength than the intact saprolitic soils and strength results of shear testing of
discontinuities are given in Tables B1.1 and B1.2 of Appendix B (after Irfan and
Woods, and Irfan). The presence of relict discontinuities affects both the hydrogeology
(discontinuities typically of lower permeability) and slope stability, depending on their
orientation in relation to the slope.
For the compound slides (Table 3.15), the inclination of the surface of rupture is 5 to
20 degrees greater than the basic friction angle. This is slightly less than the difference
obtained by Glastonbury and Fell (2000) and is possibly due to a combination of the
weaker strength of saprolitic soil mass (than intact rock strength), relatively low stresses
within the failure mass and intense hydro-geological conditions of Hong Kong.
For the translational slides the difference ranges from 5 degrees less than the basic
friction angle (Fei Tsui Road) to 15 degrees greater (Fung Wong Reservoir). For Fung
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.81

Wong Reservoir part of the rupture surface was through intact saprolitic soils and is
possibly the reason for the relatively high difference. Fei Tsui Road on the other hand
(Figure 3.42) presents an interesting example of a translational slide that developed into
a “rapid” debris slide. On first impression it is unlikely to consider that the slide could
develop into a “rapid” slide post failure given the low angle of the surface of rupture
and strong defect control on the lateral margins. It is considered that the brittleness for
this “rapid” landslide is likely to be associated with the dilation required on the irregular
surface of rupture and possibly some brittleness on the lateral margins. It is also
considered possible that altered tuff layer may have been pre-sheared to some extent
and the basic friction angle lower than given in Table 3.16. The hydro-geological
conditions present at the time of the landslide must have been sufficient to overcome the
dilation effects to result in failure of the slope.
The limited number of slides and the limited breadth of the geological environment
of the cases make it impossible to give general conclusions with any degree of
confidence. However, the principle that basal slide surfaces need to be steeper than the
residual friction angle for defect controlled slope failures to developed into “rapid”
landslides is logical, and the findings presented can be used as a guide for translational
and compound defect-controlled slides.

Approximate post-failure
surface profile

Figure 3.42: Cross section through the failure and “rapid” landslide at Fei Tsui Road,
Hong Kong (GEO 1996a)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.82

Table 3.15: Features of defect controlled compound slides that developed into “rapid” landslides from Hong Kong
Slide Name Description Slope α base Estimated Estimated Approx Rupture Surface
Angle (º) (º) φ b (º) φ r′ (º) φ b − α base Description
Ching Cheong Road (1997) Cut slope in granitic saprolite 48 40 to 50 26 to 34 18 to 22 -15 ± 5 Infilled kaolinitic joint, persistent
Allway Gardens (1993) Cut slope in volcanic saprolite and PW 41 23 to 47 22 to 30 22 -10 ± 5 10 mm thick clay infilled relict
rock joints
10000 Buddhas Monastery Cut slope in granitic saprolite and PW 52 30 to 52 35 ± 2 31 -5 ± 2 Rough, silt infilled joints
(1997) rock
Po Shan Road (1972) Cut slope in volcanic saprolite 57 43 26 ± 4 18 to 22 -20 ± 5 Persistent relict joint
Shum Wan Road (1995) Quasi-natural slope of volcanic saprolite 31 10 to 51 26 to 30 21 -5 ± 2 Clay seam of 100 to 350 mm
thickness.
α base = rupture surface inclination φ b = estimated basic friction angle φ r′ = estimated residual friction angle PW = partially weathered

Table 3.16: Features of defect controlled translational slides that developed into “rapid” landslides from Hong Kong
Slide Name Description Slope α base Estimated Estimated Approx Rupture Surface
Angle (º) (º) φ b (º) φ r′ (º) φ b − α base Description
Kau Wa Keng San Cut slope in granitic saprolite and PW 45 to 70 45 35 ± 2 22 to 30 -10 ± 5 Along relict discontinuities parallel to
Tseun, A (1997) rock the slope
Fung Wong Reservoir Cut slope in granitic saprolite 52 45 to 60 26 (defect) 20 to 25 -15 ± 10 Kaolin infilled relict discontinuity
(1998) 40 to 42 (intact) (failure partly through intact soil)
Fei Tsui Road (1995) Cut slope in volcanic saprolite and PW 60 to 65 19 24 ± 2 18 ± 4 +5 ± 2 Altered tuff layer, 0.6 m thickness,
rock (10 to 25) completely decomposed, kaolin veins.
Ka Wa Keng Upper Cut slope in granitic saprolite 60 30 25 to 30 20 to 25 -3 ± 2 Persistent 20 mm thick relict joint,
Village (1997) kaolin and maganiferous deposits.
α base = rupture surface inclination φ b = estimated basic friction angle φ r′ = estimated residual friction angle PW = partially weathered
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.83

3.6 SUMMARY OF THE CHARACTERISTICS AND M ATERIAL TYPES OF


“RAPID” LANDSLIDES
As previously discussed, it is important to distinguish between the trigger mechanism/s
for the processes leading up to slope instability and the occurrence of “rapid”
landsliding post failure. Once an unstable slope condition is reached the progression to
“rapid” sliding is dependent on the slope geometry and material properties within the
unstable slope. Whilst rainfall and hydro-geological conditions may be a significant
factor in the development of an unstable slope condition in a number of slide classes
considered, it does necessitate that the post failure movement will be “rapid”. In fact
the material properties and slope geometry are far more significant influences on the
likelihood or otherwise of “rapid” sliding.

3.6.1 FLOW SLIDES IN CONTRACTANT SOILS SUSCEPTIBLE TO FLOW LIQUEFACTION

A wide variety of soil types covering a broad spectrum of particle size distribution
(Figure 3.43) are susceptible to contraction on shearing and development of flow
sliding. For cohesionless soils Fell et al (2000) consider the material types susceptible
to liquefaction and flow sliding include submarine and sub-aqueous slopes, mine
tailings, dredged and hydraulically placed fills, dumped mine wastes, uncompacted or
poorly compacted fills, loose dumped rockfills and loose colluvium on hill slopes.
Sensitive clays derived from leached marine deposits are also susceptible to liquefaction
and flow sliding. Nearly all of these cases could be considered as dumped or loosely
placed man made fills or recent natural deposits. Being contractant in drained
conditions, slopes in these soil types are potentially susceptible to liquefaction and flow
sliding.
General guidelines on relative density of cohesionless soils susceptible to flow
liquefaction based on analysis of the literature are as follows:
• Clean sands are susceptible to flow liquefaction at a density index less than about 15
to 30%, but this is likely to vary depending on the shape of the grading curve and
particle size (i.e., coarse sands are likely to be less susceptible than fine sands). A
complete liquefaction condition (zero residual undrained strength) is observed in the
laboratory for very loose sands (relative density less than about 10%, but this will
also vary).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.84

• Silty sands are more susceptible to flow liquefaction than clean sands (in terms of
relative density) with flow liquefaction possible at relative densities up to 45 to
60%.
• Sandy gravels are less susceptible to flow liquefaction than sands. Laboratory and
field tests on coal waste and coking coal indicate these materials are susceptible to
liquefaction at void ratios greater than about 0.3. Moisture content at placement
(dumping) has a significant effect on the initial void ratio of these materials. Field
case studies indicate that flow slides in mine waste spoil piles generally occur when
the spoil pile is located on a hillside and has been placed by tipping onto the crest of
the active dump.
• Silty sands (typically with clay contents less than about 10%) are susceptible to
liquefaction at densities below 85 to 90% of Standard Maximum Dry Density (from
laboratory tests on decomposed granite in Hong Kong). Flow slides in these
materials are typically of shallow depth (up to 3 m) and therefore effective stresses
are low.

100

3
80
4
Percent passing (%)

1
2
60

40

20

0
0.001 0.01 0.1 1 10 100
Clay Silt Sand Gravel
Particle size (mm)
1 - Coarse grained coal mine waste
2 - Hydraulic placed fills (tailings and dam embankments)
3 - Sensitive clays
4 - Sub-aqueous slopes (natural and fill slopes)

Figure 3.43: Particle size distributions of material types susceptible to liquefaction and
flow sliding.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.85

These guidelines and particle size distributions are general and should be used with
caution as soil types outside these limits may liquefy and conversely soil types within
these limits may not.
Whether or not liquefaction and subsequent flow sliding will occur is dependent on
the material properties, degree of saturation and the effective stress conditions. The
mechanism for initiation of flow sliding under static conditions requires:
• The soil to be in a saturated or near saturated state.
• The effective stress state within the slope to be at or above the instability line (or
collapse surface) (Figure 3.23), under which conditions the soil is contractive on
shearing in drained loading. It should be recognised that under these stress
conditions the potentially contractile soils can exist in a stable state in drained
conditions.
• The trigger to liquefaction and flow sliding is a change in effective stress conditions,
and most likely the shear deformations that occur as a result of the change in stress
conditions.

The slope geometry, conditions under which liquefaction is triggered and factors
affecting the development of flow slides within the various soil types considered is
discussed in Sections 3.4.1 to 3.4.7.
The general guidelines (and discussion in Sections 3.4.1 to 3.4.7) provide an
indication of the potential soil types and conditions under which flow slides have
developed. They should be used with caution and as a general indicator of potential for
flow sliding. Other methods (laboratory and field based) are available and should be
used for further evaluation of liquefaction and flow slide potential.

3.6.1.1 Laboratory Based Methods

The instability line or collapse surface concept (Figure 3.7) is a useful theoretical
model, appropriate to liquefaction susceptible soils, for estimating the stress conditions
at the onset of strain weakening (and therefore liquefaction) in undrained conditions.
Soils with a positive state parameter (as determined by their void ratio) are strain
weakening in undrained loading and may be susceptible to flow liquefaction. Several
studies have attempted to define parameters for the instability line; however, its location
will vary depending on the sample void ratio.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.86

Current research queries the uniqueness of the steady state line. Recent work by
Cubrinovski and Ishihara (2000) indicates that the steady state line in void ratio versus
log effective normal stress space is not linear below effective normal stresses of 10 kPa
(Figure 3.6).
The methodology for assessment based on laboratory testing is relatively straight
forward, however, the application of theoretical concepts from laboratory testing to site
conditions have proven difficult. Most of the difficulty is associated with obtaining
“undisturbed” samples of soils, the sensitivity of the soil behaviour to small changes in
void ratio, complexity of factors such as in-situ stratification and the actual test results
themselves. If laboratory testing is to be used to assess the potential for flow
liquefaction, care must be taken to model the void ratio, fabric and initial stress
conditions of the in-situ soil, and the likely stress path to failure.

3.6.1.2 Field Testing

For assessment of flow liquefaction potential of sandy soils several useful in-situ
methods have been developed, some derived purely on an empirical basis from case
study analysis and some from a combination of theoretical and empirical analysis.
Cubrinovski and Ishihara (2000), Baziar and Dobry (1995), Ishihara (1993), Robertson
et al (1992) and Sladen and Hewitt (1989) have developed useful guidelines for
assessment of the boundary between flow and no-flow conditions of sands and silty
sands under earthquake and static loading, and in terms of SPT N value and/or CPT qc.
The most useful for static liquefaction is considered to be the recent research by
Cubrinovski and Ishihara (2000) where the flow potential takes into consideration
particle size distribution, fines content and particle angularity (Figure 3.11 and Figure
3.15), and is applicable for sands and silty sands with up to 30% fines content.
Figure 3.44 presents a comparison of the liquefaction flow no-flow boundaries in
terms of SPT (N1)60 value derived from field cases of earthquake induced flow
liquefaction (Baziar and Dobry 1995; Ishihara 1993) and derived from (with some
exceptions) laboratory undrained monotonic triaxial compression tests in sands and silty
sands (Ishihara 1993; Cubrinovski and Ishihara 2000). The upper bounds of Ishihara
(1993) and the bounds by Cubrinovski and Ishihara (2000) have been converted from
SPT N value to (N1)60 correcting for over-burden pressure and energy ratio (assuming
78% rod energy for the SPT results from Japan). The liquefaction flow no-flow
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.87

boundaries of Baziar and Dobry (1995) and Ishihara (1993) compare reasonably well
with those of Cubrinovski and Ishihara (2000) in consideration of the materials they
represent (Ishihara (1993) upper bound for clean sands and silty sands (up to 30%
fines), Baziar and Dobry (1995) sands with at least 10% silt).
Based on the comparisons between field and laboratory, and between static and
cyclic trigger (Figure 3.44 for SPT and Figure 3.10b for CPT), it is concluded that the
flow liquefaction no-flow boundary is dependent primarily on the properties of the
material and is independent of the triggering mechanism; i.e., cyclic (earthquake) or
static. Therefore, importantly, the methodology of Cubrinovski and Ishihara (2000) can
be applied to the assessment of flow liquefaction potential for clean sands and silty
sands (up to 30% fines) for static and dynamic triggered events.

Figure 3.44: Comparison of flow liquefaction no-flow boundaries (in terms of SPT
(N1)60) for sands and silty sands from monotonic laboratory undrained tests and
earthquake triggered field cases (after Cubrinovski and Ishihara 2000).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.88

3.6.2 CHARACTERISTICS AND IDENTIFICATION OF CONDITIONS UNDER WHICH


FAILURES IN DILATIVE SOILS RESULT IN “RAPID ” LANDSLIDING

The “rapid” landslides developed from failures in dilative soils were divided into two
categories, those where the surface of rupture passed through the soil mass, and those
where it is preferentially located along defects within the slide mass, termed defect
controlled slides. A greater weighting was given to the slides through the soil mass due
to the greater weighting of case study information. An important aspect of landslides
from dilative soils is the distinction between the initial development of the failure itself
and the transformation of the slide mass into a “rapid” debris flow or debris slide.
Initiation of the failure occurs when a critical equilibrium condition is reached within
the slope along a defined surface of rupture and is dependent on material strength, slope
geometry and the hydro-geological conditions within the slope.
For slides through the soil mass in natural slopes the significant factors affecting
initiation of the slide itself and its transformation into a “rapid” debris slide or debris
flow are:
• Rainfall is often the trigger for initiation of the slide. Correlations of rainfall to
landsliding, while useful, are not straightforward due to the influence of the
hydrogeology, material properties and slope geometry on slope stability.
• Material properties and particle size distribution. The range of particle size
distribution and material properties is very broad, ranging from high plasticity silts
to low plasticity sandy clays and clayey silty sands to dominantly coarse granular
soils with low fines contents (less than about 10 percent). Figure 3.40 and Table
3.12 provide guidelines on material properties.
• Slope geometry. Landsliding predominantly occurs where the source and
immediate down-slope angle is greater than about 25 degrees, but can occur on
slopes down to about 18 to 20 degrees. The slides occurred on open slopes as well
as in areas conducive to concentration of water (concavities, breaks in the slope or
geological contacts).

Two modes of post failure deformation and slide mass behaviour are evident for
slides through the soil mass in natural slopes:
• Slides of debris that remain relatively intact during travel (i.e. debris slides). The
higher clay content soil types (high plasticity silts and low to medium plasticity
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.89

sandy clays and clayey sands) generally remain relatively intact during travel and
reach maximum slide velocities in the rapid to low end of very rapid category.
• Slides of debris that undergo significant break up and remoulding, transforming into
debris flows. Material types typically range from clayey silty sands to coarse
grained granular soils (predominantly cobble to boulder size) with low fines
contents. Travel velocities are much higher and typically in the range very rapid to
extremely rapid.

For slides through the soil mass in cut slopes the case studies are from one region
(Hong Kong) and occurred in colluvium, residual soils and decomposed rock types
derived from granitic and volcanic rocks. Material types were generally silty sands to
sandy silts with low clay contents containing varying quantities of gravel to boulder
sized material. In Hong Kong, “rapid” landslides developed from failures in cut slopes
that were greater than about 34 degrees. Studies in other soil types suggest sliding can
occur in cut slopes as low as 27 to 30 degrees, but is dependent on the strength
properties of the soil and the location of the slide within the slope.
For the defect controlled slides in dilative soils only a limited number of slides were
analysed and they were all from a similar geological environment, making it impossible
to give general conclusions with any degree of confidence. However, the analysis
indicated that “rapid” sliding occurred where the inclination of surface of rupture is:
• 5 to 20 degrees greater than the basic friction angle for compound slides.
• Equal to or greater than the basic friction angle for translational slides.

The principle that basal slide surface need to be steeper than the residual friction
angle for “rapid” sliding to occur post failure is a logical one, and the findings can be
used as a guide for translational and compound defect-controlled slides.

3.7 P RE- FAILURE DEFORMATION BEHAVIOUR


The pre failure deforma tion behaviour can give warning of an impending failure, and
allow evacuation of persons, closure of roads, etc., so mitigating risks. The following
summarises the observations of pre failure deformation made for each class of slope.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.90

3.7.1 FLOW SLIDES

Monitoring records of pre-failure deformation of flow slides has been obtained for a
limited number of case studies. Most of these are from coal waste stockpiles in British
Colombia where monitoring of deformations at the crest of the tip is used as a control
measure in an attempt to avoid a potential failure by halting of waste deposition and
also as a measure to reduce the risk to loss of life and property. Monitoring of
deformation was also recorded for the two flow slides in embankments during
construction (Fort Peck Dam and Calaveras Dam). For the remainder, records of pre-
failure deformation are observational. In most cases no formal monitoring is (or was)
undertaken.
Discussion and actual records of pre-failure monitoring for the various sub-groups
of flow slides is given in Sections 3.7.1.1 to 3.7.1.4.

3.7.1.1 Coarse-Grained Coal Mine Waste Spoil Piles on Hillsides and Coking Coal
Stockpiles
Pre-failure deformation was measured or observed at a number of the failures in coarse-
grained dumped fills at British Columbia, South Wales and Hay Point. The records and
observations are summarised in Table 3.17. Recorded measurements of crest
displacement leading up to failure at two waste dumps in British Columbia are shown in
Figure 3.45.
Where recorded or closely observed, the rate of displacement leading up to an
impending failure increased significantly in the days and sometimes weeks prior to the
failure reaching, maximum rates of 2.5 to 50 m/day in the last 1 to 2 hours prior to
failure and “rapid” movement of the slide mass. This observation of increasing rate of
deformation prior to failure is quite common in monitored failures in soil and rock
slopes.
Hungr and Kent (1995) comment that most of the coal mines in British Columbia
have criteria for closure of dumping operations based on the rate of deformation.
Monitoring of the coal waste spoil piles (in British Columbia) indicates that
deformations during active tipping, even if of large magnitude, and tension cracking do
not necessarily indicate an impending failure condition. The deformation behaviour of
the coal mine waste spoil piles (in British Columbia) that was considered as “normal”
behaviour are:
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.91

• Rates of deformation during active tipping up to 1 to 1.5 m/day are not uncommon
and do not necessarily indicate an impending failure condition (Hungr and Kent
1995; Campbell and Shaw 1978).
• Total crest displacements in excess of 9.1 m have occurred that did not lead to a
slope failure (Campbell and Shaw 1978).
• Tension cracking at the crest is also a common feature of the coal waste dumps and
is not an indication of impending instability (Nichols 1982).
• Foundation slope has a significant effect on the rate of deformation whilst
precipitation has a negligible effect (Nichols 1982; Campbell and Shaw 1978).
Monitored rates of deformation at one mine decreased significantly as the toe of the
waste pile advanced onto a flatter slope (reduction in rate from average of about 0.7
m/day to less than 0.2 m/day as the toe advanced over a 27 degree slope to a 14
degree slope).

Deformation monitoring of coal mine waste spoil piles in South Wales and coking
coal stockpiles at Hay Point, Australia is (or was) not undertaken. However,
observations of failures indicate that “abnormally” large deformations and high
deformation rates preceded the failure. The observations from South Wales, although
similar in some ways to British Columbia, comme nt on the development of a distinctive
“boot” shape or bulge at the toe of the spoil pile of failed slopes.

3.7.1.2 Hydraulic Fill Embankment Dams

Pre-failure deformation was observed at Fort Peck and Calaveras Dam. In both cases
the failure occurred during construction. At Calaveras Dam, horizontal deformations
had been recorded for some nine months prior to the failure and totalled some 1100 mm
up to the day prior to failure. On 18 June 1917, some 600 mm of deformation occurred
in 1.5 days and sluicing operations were shutdown for a period of time. After this date,
further deformations occurred during periods of sluicing in the nine months leading up
to the flow slide on 24 March 1918.
At Fort Peck Dam initial deformations were a result of shear displacements on
slickensided bentonitic seams within the Bearpaw Shale formation underlying the
alluvial, predominantly sand, embankment foundation. Over-night the upstream bank
of the core pool had settled some 450 to 600 mm and continued up to the time of failure
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.92

and flow sliding at 1:15 pm that same day. Eyewitness reports (ENR 1939a) indicate
the core pool began to settle slowly at first, and then at an increasing rate.
In both cases large deformations at an increasing rate of deformation were observed
prior to the failure.

Table 3.17: Summary of pre-failure observations in coarse-grained loose fills


Case Study Rate Observation / Comment Reference
(m/day)
British Columbia 30 (max) Total cumulative displacements exceeding Hungr and Kent
(B.C.), general 3.5 (1 day prior) 10s of metres and rates up to 1 m/day do not (1995)
indicate failure.
Impending failure indicated by acceleration of
crest displacement to rates of 1 to 5 m/day (up
to 1 m/hour prior to failure).
Fording, B.C. 15 (hour prior) Total displacement of about 8.5 m over period Campbell and Shaw
(22/6/72) 1 (day prior) of 3 weeks (refer Figure 3.45). (1978)
Fording, B.C. 2.5 (hour prior) Total displacement of about 8.2 m over period Campbell and Shaw
(2/7/72) of 6 days (refer Figure 3.45). (1978)
Fording, B.C. - Recorded rates up to 1.5 m/day and total crest Campbell and Shaw
displacements in excess of 9 m have not (1978)
resulted in failure.
Quintette 1660, - Large cracks several hours prior to failure. Dawson et al (1998)
B.C. (9/9/85) Major cracking and bulging not evident until
just before failure.
Greenhills Cougar - Cracks observed in dump face Dawson et al (1998)
7, B.C. (11/5/92)
Fording South, 25 to 30 Dump closed due to high rates of deformation Dawson et al (1998)
B.C. (29/10/89) (immed. prior) on 24/10/89, 5 days prior to failure.
2.4 (day prior)
Aberfan, Tip 7, Sth 30 to 50 Approx. 6 m crest settlement in 12 hours prior Bishop et al (1969)
Wales (21/10/66) (immed. prior) to the failure, 3 m in the last 1.5 to 2 hours.
Deformations observed 3 to 6 months prior to
failure.
Aberfan, Tip 4, Sth - Bulging observed in months prior to failure. Bishop et al (1969)
Wales (21/11/44) Some minor sliding before final failure.
Cilfynydd, Sth - Deformation observed on day prior to failure. Bishop et al (1969)
Wales (5/12/39) Survey suggests toe bulging well before
failure.
South Wales - Toe bulging a common observation of failed Siddle et al (1996)
spoil piles. Pre-failure deformation observed
at Maerdy (1911), Mynydd Corrwg Fechan
(1963) and Pentre (1909).
Peak Downs, Hay - Longitudinal cracking observed on face of Eckersley (1985)
Pt., Aust. (3/2/77) stockpile prior to failure.
B.C. = British Columbia, Canada
Hay Pt., Aust. = Hay Point, Australia
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.93

Figure 3.45: Pre-failure deformation of two separate failures at the Clode waste dump,
Fording Coal, British Columbia (Campbell and Shaw 1978).

3.7.1.3 Sensitive Clays


As discussed in Section 3.4.3 retrogressive flow slides in sensitive clays generally
develop from an initial failure, usually within a steep river or stream bank subject to
active erosion. The pre failure deformation behaviour of interest is therefore associated
with this initial slope failure.
Pre-failure deformation has been recorded for a small number of natural and man-
made slopes in Canada (Mitchell and Eden 1972; Eden 1977; Mitchell and Williams
1981). The measured rates of deformation from six slopes adjacent to slopes that failed
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.94

are presented in Table 3.18. The deformation behaviour from these slopes is considered
to be representative of deformation rates of the initial slide in similar slopes (and soil
types) that can develop into large-scale retrogressive landslides. Tavenas (1984)
comments that the increased rates of deformation observed in spring are associated with
higher groundwater conditions following snowmelt and seasonal rainfall and therefore
at effective stress states closer to a failure condition (i.e. at a reduced factor of safety).
The typical deformation profile within an instrumented slope (Figure 3.46) indicates
the pre-failure deformation has developed within the soil mass rather than along a
distinct shear band and also shows some rotation of the inclinometer tube. The analysis
of failures in embankments on soft ground (Chapter 4) shows that a similar deformation
profile is observed whilst embankment construction is in progress, and, for sensitive
clay foundations, it is not until shortly prior to slope failure that shearing within a
localised shear band is observed.
The recorded slope deformations (Figure 3.47) from an instrumented natural slope
brought to failure by raising groundwater levels through recharge wells (Mitchell and
Williams 1981) indicates that the rate of deformation is affected by the stress state
within the slope. The onset to failure occurs several days prior to slope failure when the
pore pressure ratio is increased from 0.43 to 0.47 and the rate of deformation increases
markedly.

Figure 3.46: Measured deformations in a natural slope, Ottawa River (Mitchell and
Eden 1972)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.95

Table 3.18: Measured slope deformations adjacent to failures in the Ottawa area (Mitchell and Eden 1972)
Site of Inclinometer Monitoring Site of Recent Landslide
Location Description Slope *1 Rate of Deformation Time of Description of failure Slope *1
(mm/year) Maximum
Height Angle 3 year Maximum Deformation Height Angle
(m) (deg.) Average Rate (m) (deg.)
Rockcliffe, R1 Natural slope (typical for bank) 14 26 2.5 25 Spring 6000 m3, retrogressive, 1969 13 28
3
Green Creek, G1 Natural slope (typical), no toe 19 28 1.1 35 Spring 6000 m , retrogressive, 1971 20 26
erosion
Breckenridge, B1 Natural slope, toe loading from 29 27 Not known 70 Spring 23000 m3, retrogressive, 1963 28 25
recent landslide
Rockcliffe, R2 Graded natural slope 13 20 13 700 After slope 23000 m3, retrogressive, 1967 13 28
grading
Green Creek, G2 Steep natural slope 15 35 2.7 25 Spring 1500 m3, 1969 (monolithic?) 19 26
3
Breckenridge, B2 Natural slope, subject to rapid 28.5 22 - 500 Autumn 23000 m , retrogressive, 1963 28 25
erosion. The slope failed.
Note: *1 Slope angles are an average over the height of the slope.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.96

Figure 3.47: Deformation and pore pressure ratio of an induced slope failure (Mitchell
and Williams 1981)

3.7.1.4 Flow Slides in Other Soil Types

No published records of pre-failure deformation prior to “rapid” landsliding could be


found for flow slides in loose sandy and fine-grained fills (tailings dams, loose fills
from Hong Kong) or natural slopes (submarine slopes, loessic soils and other colluvial
soils). Based on the pre-failure deformation and failure mechanism from flow slides in
other soil types, it is considered that:
• For flow slides that develop from an initial, essentially drained type failure within a
slope of limiting stability, pre-failure deformation would be observable if monitored.
This would apply to several failures in tailings dams and failures in submarine
slopes where the initial failure occurred in the steep outer slope and was triggered by
changes in effective stress. The initial failures of the tailings dams at Stava,
Arturus, Saaiplass and Buffalo Creek were considered to be drained type failures of
the outer slope triggered by high perched or groundwater conditions. Previous slope
failure (Buffalo Creek) and toe sloughing (Merriespruit) confirmed the limiting
stability of these slopes.
• For flow slides triggered by intense or heavy rainfall (loose fills in Hong Kong, and
loessic and other colluvial soils) some, but very limited, pre-failure deformation
would be expected due to changes in effective stress conditions during the rainfall
period. During the rainfall event, leading up to failure, the stability of a slope would
be expected to reduce to one of meta-stability due to the reduction in matrix suction
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.97

of the partially saturated soil and the development of perched water tables in these
hydro-geologically complex slope profiles.

Laboratory undrained triaxial compression tests on loose fill materials derived from
completely decomposed granite (HKIE 1998) indicates peak strengths were reached at
small axial strains (1 to 2%) prior to the rapid onset of static liquefaction. This would
indicate that limited pre-failure deformation occurs prior to failure in these loose fills.
However, it is recognised that the effective stress path in these laboratory tests is not
representative of the actual effective stress path that would be observed in the field.

3.7.2 “RAPID ” LANDSLIDES FROM SLOPE FAILURES IN DILATIVE SOILS

3.7.2.1 Slides Through Soil Mass – Dilative Failures

No published records of pre-failure deformation prior to slope failure could be found for
“rapid” landslides that developed from failures in dilative soil types; including the slides
in natural and cut slopes in colluvial and residual soil profiles.
These slides are often triggered by intense or heavy rainfall. During the rainfall
event the stability of a slope would be expected to reduce to one of meta-stability due to
the reduction in matrix suction of the partially saturated soil and the development of
perched water tables in these hydro-geologically complex slope profiles. Pre failure
deformation would be expected to occur for these shallow dilatant failures due to
changes in effective stress conditions; however, the amount of deformation is likely to
be small.
These slides, particularly the shallow slides, generally occur during the period of
rainfall. Rainfall based systems are used to warn of the potential for landslides in
populated regions where the climate and topography make them susceptible to these
types of failures (such as Hong Kong and Rio de Janiero). Due to high incidence and
the wide spread occurrence of rainfall triggered landslides in these regions such systems
are more appropriate than methods to monitor deformation.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.98

3.7.2.2 Defect Controlled Slides


The case studies of defect-controlled slides (all from Hong Kong) occurred in saprolitic
soils derived from the weathering of granite and volcanics (typically tuff). No formal
deformation instrumentation was installed for monitoring of pre failure deformation
behaviour. However, observations at several sites provide some indication that pre-
failure deformation did occur. The observations were:
• Previous landsliding within or nearby to the reported failure in the months leading
up to the reported landslide. This was observed in 3 of the 9 cases (Po Shan Road,
1972; Ching Cheong Road, 1997; Ka Wa Keng San Tseun (Slide A), 1997).
• Evidence of open joints or cracking that existed prior to the reported failure. This
was observed in four cases (Po Shan Road 1972, Ching Cheong Road 1997, Ten
Thousand Buddas Monastery 1997 and Allway Gardens 1993). All four were
defect-controlled compound type landslides.
• Deformation and cracking in the days and weeks leading up to failure and “rapid”
movement of the failed slide mass. This was observed at one site only (Po Shan
Road, 1972) of the cases analysed. Pre-failure deformation was also observed at
Cut CC2 at Ching Cheong Road in 1972 (50000 m3 defect controlled compound
slide).

The Po Shan Road landslide of 1972 (Vail 1972) comprised the failure of a cut slope
in volcanic saprolite and a significant portion of the natural slope above the cut. The
estimated volume of the landslide was 40000 m3. It was a relatively deep failure for
Hong Kong (6 to 16 m depth) and the surface of rupture indicates a compound slide
consisting of a significant defect-controlled translational component with buttressing at
the toe. The collapse of the slope was preceded by observations of deformation
(cracking and sliding) in the nine months prior to the failure. In the days leading up to
the failure cracking and deformations were observed on and above the cut slope.
Although the deformation records are somewhat haphazard, it is considered that they
indicate an increase in the rate of deformation leading up to the slope failure.
Glastonbury and Fell (2000) concluded that the defect-controlled translational and
compound failures in rock slopes, that developed into “rapid” landslides, generally
showed precursory warning of the impending collapse, mainly in the form of cracking,
localised failures and increased noise, with greater warning for the compound slides.
Pre failure deformation monitoring records were limited but indicated an increasing rate
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.99

of deformation generally over a period of days to months prior to collapse. They found
a general trend of increasing pre collapse deformation with increasing failure volume.
The observational data for the defect-controlled slides in saprolitic soils are similar
to those observed for the rock slides, even though the slide volumes are several orders
of magnitude less. This is to be expected given the similar mechanism/s associated with
failure. The findings also indicate that the warning signs are more prevalent for the
compound slides.

3.7.3 SUMMARY OF PRE FAILURE DEFORMATION BEHAVIOUR

Pre failure deformation monitoring records of slope failures is limited to a handful of


case studies of “rapid” post failure movement. Observational records of deformation
and other pre failure warning signs are available for a larger group of the landslides.
Overall, pre failure deformation is considered to occur for all slides prior to failure
in the form of acceleration in the rate of deformation. For a number of slide groups
appropriately located deformation instrumentation is likely to provide sufficient
warning of an impending failure. However, for a number of slide groups the total
deformation associated with the onset to failure is likely to be small and the increase in
rate of deformation obscured by deformation rate changes associated with changes in
effective stress conditions. The characteristics of pre-failure deformation of slope
failures that developed into “rapid” landslides are summarised in Table 3.19, and in
summary:
• Slides through the soil mass triggered by intense rainfall were mostly of small
volume and/or shallow slides. Pre-failure deformation is limited and likely to occur
within the rainfall period. Rainfall based warning systems are useful in regions
where these slide types occur.
• For compound type defect controlled slides of slide volumes greater than about 1000
to 2000 m3, observational warning signs should be evident. Deformation
monitoring should provide adequate warning of the impending failure.
• For translational type defect controlled slides, limited observational signs of the
failure case studies suggests deformation monitoring may not provide adequate
warning of the impending failure. However, there were only four case studies in the
database (ranging in volume from 80 to 14000 m3).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.100

Table 3.19: Summary of characteristics of pre-failure deformation of slope failures that


developed into “rapid” landslides
Initial Slide Slide Total Deformation Comments
Classification Sub-Group Deformation Rates
Large, For larger volume slides significant
Coarse-grained in the order of High warning in terms of days to failure.
stockpiles metres for large (2.5 to 30 m/day) Limited information on smaller volume
volume slides slides from Hay Point.
Hydraulic fill At least 0.5 to 1 day period of
embankment “abnormal” deformation prior to
Large, up to 1 m Medium failure. Sudden and rapid increase in
dams (static
the rate of deformation at failure.
trigger)
Initial drained failure shows typical
Flow Slides in pre-failure increase in rate of
Contractile soils Sensitive clays Very low
Small deformation to failure. Small
(up to 1m/year ?)
deformations and very low rates of
deformation.
No records. Observations at failure
Tailings dams Not known Not known
were of piping induced failures.
Hong Kong failures triggered by loss of
Loose sandy suction and/or transient pore pressures
Suspect very Suspect very
fills (sub- during intense rainstorms. No
limited low
aerial) observations, suspect very limited
deformation.
Typically shallow slides in colluvial
and residual soils, and predominantly
in natural and cut slopes.
Triggered by reduction in matrix
Slides through suction and/or transient pore pressures
Suspect limited Suspect low
the soil mass due to rainfall, snowmelt. No
observations, suspect limited
deformation.
Slides in
Rainfall based systems provide better
Dilative soils
warning.
Not known. Compound slides show greater warning
Suspect moderate signs than trans lational slides.
Defect to large (up to For compound slides the typical
controlled metres ?) Not known warning signs are cracking, opening of
slides Dependency on joints, previous sliding and acceleration
slide volume and in deformation in the days and weeks
slide type. leading up to failure.

For flow slides in contractile soils it is concluded that:


• Pre-failure deformation is a pre-cursor to flow sliding in most cases where the
trigger is due to static changes in effective stress (as opposed to dynamic trigger
such as earthquake or blasting). The exceptions to this would be piping or over-
topping induced failures of tailings dam embankments.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.101

• The on-set to failure is evidenced by an increase in the rate of pre-failure


deformation above that which could normally be expected due to increasing load
(from active placement or construction) or from other changes in effective stress
conditions such as increased pore water pressures.
• The rate of shear straining (and total deformation) required to induce undrained
conditions, and therefore static liquefaction and flow sliding in undrained strain
weakening soils, is strongly dependent on the permeability of the potentially
liquefiable soil. Soils of higher permeability requiring greater rates of shear strain to
induce undrained conditions (i.e., sandy gravel compared to a sensitive clay or silty
sand).
• The undrained brittleness index, amount of potential volume contraction on
shearing, size and complexity of the landslide will also affect the pre-failure rate of
shear strain and total deformation.

With respect to the last point, the undrained brittleness index affects the rate of
progression of the liquefied zone once liquefaction has been initiated along a portion of
the surface of rupture. If the undrained brittleness index is small then the rate of
progression of the liquefied zone will be slower. Complexity of the slide is the
consideration of factors such as internal shearing for kinematic admissibility of the
slide, three-dimensional effects, layering effects and variation in soil properties. The
complexity of the slide mass is considered likely to increase with increasing volume,
and the amount of pre-failure deformation increases with complexity (e.g., failure of
Fort Peck Dam compared to Wachusett Dam, and Hay Point failures compared to say
Aberfan). The amount of potential volume contraction (related to the relative density
and undrained brittleness index) will affect the volume of pore fluid to be dissipated if
drained conditions, and stability, are to be maintained. Soils with high volume
contraction on shearing generate a greater volume of fluid to dissipate, which can lead
to very rapid progression of a liquefaction condition on the surface of rupture.
For flow slides therefore, pre failure deformation and other warning signs of
impending failure are more evident for the coarser material types within which
liquefaction occurs (e.g. sandy gravels of the coal mine waste spoil piles) and the larger
volume slides (excluding the retrogressive slides).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.102

3.8 P OST-F AILURE D EFORMATION BEHAVIOUR – REVIEW OF THE


LITERATURE

3.8.1 FACTORS AFFECTING TRAVEL DISTANCE AND VELOCITY

The mechanics at failure of flow slides and defect-controlled landslides can give the
sliding mass significant kinetic energy on failure and these slides can therefore develop
high velocities in short distances from the slide source. As indicated by Hutchinson
(1988), the degree of brittleness associated with the failure is a significant controlling
factor in the generally high velocity of post failure movement of flow slides.
Locat and Leroueil (1997) comment that a logical approach to post failure behaviour
would be to consider the energy balance in the post-failure stage, the basis of which is
used in numerical models such as those developed by Hungr (1995). The components
of the energy equation, as stated by Locat and Leroueil (1997), are:

E p = E f + Ed + Ev + Ek (3.8)

where Ep, Ef, Ed, Ev and Ek are the potential, friction, disaggregation or fragmentation,
viscous and kinetic energy respectively. However, it is difficult to estimate all these.
For flow slides in strongly strain-weakening soils (i.e. with high undrained brittleness
index) the frictional energy on the surface of rupture is significantly reduced once
liquefaction has occurred leaving the failing mass with significant kinetic energy to
quickly accelerate to very high velocities and therefore have the potential to reach very
large travel distances.
Observations of the initiation of debris slides and debris flows on steep slopes
indicate that these landslides also develop significant kinetic energy in the early stages
of deformation and therefore quickly accelerate to reach relatively high velocities. In
these cases the initial landslide is transformed into a “rapid” debris slide or debris flow
due to strength reduction associated with incorporation of water, dilation, loss of
cementing and remoulding, or due to a second phase of contractive shear of the
loosened slide mass. For soil types where the strength is sensitive to small changes in
moisture content (cohesionless and low cohesion materials) and it is possible for water
to become readily entrained in the dilating soil mass, the transformation to a “rapid”
debris slide or debris flow is generally observed to occur rapidly.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.103

Once the landslide (or part thereof) has failed, the factors affecting travel distance
and velocity of the slide mass are:
• Down-slope geometry, including the down-slope angle, changes in down-slope
angle and degree of confinement. It will be shown, particularly for cut slopes
(Section 3.9.3), sharp changes in down-slope angle (from steep to flat) dissipate the
energy of the sliding mass thereby decreasing the velocity and travel distance,
although scale effects are important.
• Confined debris flows tend to develop greater travel distance than those on
unconfined or open slopes at a similar down-slope angle (Corominas 1996a; Golder
Assoc. 1995).
• The presence of water. Water within the travel path will generally become entrained
in the slide mass resulting in overall reduction in strength and greater travel
distance. For mine tailings dams (Section 3.9.4.3 and 3.10.3) the presence of water
can alter the flow behaviour.
• Volume of slide mass. Analysis of “rapid” slides (Corominas 1996a; Hutchinson
1988) indicates that the greater the slide volume the greater the travel distance (by
virtue of the decrease in travel distance angle with increase slide volume).
• Material type of the slide mass. Cohesionless (or low cohesion) materials are more
prone than cohesive materials to strength loss on entrainment of water (Johnson and
Rodine 1984). Conversely, the rate of dissipation of pore pressures developed
during flow or on initiation (flow slides) would be significantly lower in soils with
high fines contents (finer than 75 micron), such as for sensitive clays compared to
sandy gravels.
• Material mantling the down-slope. Dawson et al (1998) and Golder Assoc. (1995)
comment that the travel distance of flow slides from coal waste stockpiles increased
when the foundation soils were susceptible to strength loss on rapid loading from
the slide mass. The soil types susceptible (Golder Assoc. 1995) were saturated (or
near saturated), loose, finer grained soil deposits, particularly those with a high
proportion of organic matter; i.e., those soils of relatively low permeability within
which high pore water pressures can develop due to contraction on rapid (and high
impact) loading. This excluded soils with a low degree of saturation (high pore
pressures cannot be generated), high relative density (dilative under high impact
loading) and coarse granular grading (i.e. free draining).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.104

• Obstructions to the slide mass dissipate energy thereby decreasing the velocity and
travel distance (Corominas 1996a).

For any specific group of “rapid” landslides (e.g. from dilatant slides in cut slopes,
from flow slides in mine tailings dams, etc.), the dominant factors influencing the travel
distance and velocity will vary. For flow slides in Canadian mine waste spoil piles
Golder Assoc. (1995) concluded that the dominant influences on travel distance were
the degree of confinement of the slide mass and the presence of potentially liquefiable
materials mantling the down-slope. For flow slides in mine tailings dams the down-
slope angle and presence of free water have a dominant influence. Where the down-
slope angle is relatively flat (for flow slides in mine tailings), the flow is typically
laminar; however, for down-slope angles steeper than about 5 to 7 degrees the flow
behaviour of fluid tailings will become turbulent. Laboratory results (Blight 1997)
indicate that the shear strength of the tailings is strongly dependent on the moisture
content (Figure 3.83) and if water is present on the down-slope the basal portion of the
flow (for laminar flow on relatively flat down-slope) will entrain the free water thereby
reducing the shear strength and result in much greater travel distance.
There is also significant co-dependence between the factors affecting travel
distance. For the post-failure behaviour of Canadian waste spoil piles Golder Assoc.
(1995) the potentially liquefiable materials were generally located in confined valleys
and therefore the very long travel distance events with low travel distance angle were
for a group of confined flow slides where liquefiable materials were present in the gully
region.

3.8.2 VELOCITY OF THE SLIDE MASS

Reported velocities of the slide mass for debris slides, debris flows and flow slides in
soil slopes are summarised in Table 3.20, for the case studies analysed, and Table 3.21,
for debris slides and debris flows from Corominas et al (1996) and Corominas (1996b).
All rates of deformation are within the extremely rapid and very rapid classifications for
landslides (IUGS 1995, refer Table 1.1).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.105

Table 3.20: Velocity of the slide mass for “rapid” landslides from failures in soil slopes
(mostly from case studies analysed).
Location Type of Slide Velocity *1 Down-Slope Comments
(travel classification) (m/sec) Angle
(degrees)
Debris slides and debris flows (generally) 0.3 to 30 various Refer Table 3.21 for
sources.
Hong Kong Flow slides in loose fill 5 to 15 18 to 30 DAN model (Hungr
Debris flows 8 to 15 GR* 2 1998; Ayotte and
Confined debris flows 9 to 24 Hungr 1998)
Coal waste spoil Flow slide typically 15 to various sources
piles - British - Confined 30 to 60 (D), 25 (initially)
Columbia 4 to 27 (A)
- Unconfined 20 to 40 (D)
Coal waste spoil Flow slide 3 to 10 10 to 13 Bishop (1973), Siddle et
piles - South Wales al (1996), Bishop et al
(1969)
Hydraulic Fill Flow slide 0.6 to 9 -2 to 2 Various sources
Embankment dams
Tailings dams Flow slide 2 to 25 0 to 8 Various sources
Quick clays Flow slide 0.2 to 15 Not known Various sources
Submarine slopes Flow slide 0.8 to 8 Not known Average rates. Edgars
and Karlsrud (1983) and
other sources.
Notes: Refer to summary tables in Section 1 of Appendix B for sources of information on specific slides
*1 D = DAN model (Golder Associates 1995), A = actual
*2 Hungr GR (1998) = Hungr Geotechnical Research (1998)

Two methods are available for prediction of velocity; they are prediction from simple
sliding block models and from numerical continuum models.
The sliding block models, summarised by Fell et al (2000), analyse the sliding mass
as a dimensionless lumped mass moving from the centre of gravity of the pre-failure
position to the centre of gravity of the post slide position (Figure 3.48). The line
connecting these two centres of gravity represents the kinetic energy head and its
alignment is dependent on the type of rheological model used to model the movement of
the sliding mass (Figure 3.48). The kinetic energy of the lumped mass at any point is
represented by the vertical elevation of the kinetic energy head line above the slide path.
Assuming the resistance to motion is purely frictional (therefore the line joining the
pre-failure and post slide centre of gravities will be a straight line) the kinetic energy of
the slide mass equals v 2 / 2 g , where v is the velocity of the centre of the slide mass, and
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.106

the velocity can be estimated. This method presents a relatively simple means of
estimating the velocity, which does not properly model the loss of strength on
disaggregation of the slide mass or the deposition/accumulation on the travel path.
Simple sliding block models, tends to predict velocities higher than the actual estimated
velocity.

Table 3.21: Velocity of slide mass for debris slides and debris flows
Reference Slide Type Velocity Comments
(m/s)
San Francisco Bay region – Debris flows in 8 to 22 High velocities in granular soil
Ellen et al (1988), Wieczorek et sandy soils types. Much lower velocities for
al (1988), Howard et al (1988) Debris slides in metres/min medium to high plasticity silts,
clayey soils sandy clays and clayey sands
Hutchinson (1988) Debris flows and Max. 30 Velocity increases with increasing
flow slides slope angle and decreases with
increasing clay content
Johnson and Rodine (1984) Debris flows 3.5 to 11 Source area slope angle 20 to 45
degrees
Pierson and Costa (1987) Dry and wet debris 0.1 to 35
flows
1
Terzaghi (1950) * Debris flows 3 to 4.5
Curry (1966) *1 Debris flows < 16
1
Johnson and Hampton (1969)* Debris flows 3.5 to 12
1
Morton (1971) * Debris flows 0.5 to 3.5
Campbell (1975) *1 Debris flows 0.3 to 12 Source area slope angle 22 to 45
degrees
Sauret (1987) *1 Debris flows 0.5 to 15
Bovis (1993) *1 Debris flows 5 to 10
1
Lowe (1993) * Debris flows 0.3 to 12
Corominas (1996b) Debris slides and < 16 Initiating from “slides of debris”
debris flows
Note: *1 sourced from Corominas et al (1996), mostly confined debris flows

Figure 3.48: Sliding block model showing kinetic energy lines for different rheological
models (Golder Assoc. 1995).
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.107

Better methods of prediction of velocity use numerical continuum models that have
been calibrated against historical events of “rapid” sliding in similar soil types with
similar morphological, topographical and climatic conditions within a consistent
rheological framework that can then be applied to prediction of identified potential
landslides.
The results of numerical continuum modelling (by Hungr in association with others)
using the DAN program for “rapid” landslides from slope failures in Hong Kong and
coal mine waste spoil piles in British Columbia are discussed in Section 3.10. They
found that:
• The frictional model gives reasonable predictions of velocity for slide masses
travelling on open and partly confined travel paths.
• For confined travel paths the frictional model predicts much higher velocities than
estimated velocities of the actual slide. The Voellmy model (frictional – turbulent
model) provides better predictions of velocity for confined travel paths.

3.8.3 EMPIRICAL METHODS FOR ASSESSMENT OF TRAVEL DISTANCE

The available empirical methods for assessment of travel distance originated from work
on very large landslides by Heim (1932), Scheidegger (1973) and Abele (1974) who
recognised that the travel distance angle of the slide debris reduced with increasing
volume. Empirical equations of the form of Equation 3.11, derived from historic
events, could be used to predict the travel distance angle (Figure 3.49) of a potential
landslide based on estimation of the slide volume. The travel distance could then be
predicted by application of the travel distance angle to the slope geometry (Figure 3.49)
and the velocity could then be assessed from the kinetic energy of the slide mass (Figure
3.48). Hsü (1975), Sassa (1988), Van Gassen and Cruden (1989), Hutchinson (1986),
Koerner (1977), amongst others, further developed the understanding of travel
behaviour of the slide mass.
Empirical correlations (Equation 3.9) have also been developed for depositional area
to slide volume for rock avalanches (Li-Tianchi 1983; Hungr 1990; Smith and Hungr
1992). Hungr (1990) used this approach, with the assumption of a simplified bi-linear
travel profile (after Hsü 1975), for prediction of the excess travel distance, Le (Figure
3.49), in the form of Equation 3.10. A review of these area to volume concepts shows
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.108

they seem to present no advantages over other methods (discussed further in Section
3.9.4.2).

A = cV n (3.9)

(
Le = R ∗ c ∗ V n )
0.5
(3.10)

where A = depositional area (in square me tres), V = slide volume (in cubic metres), c
and n are constants, and R, the spreading ratio, is a measure of the degree of lateral
spreading of the deposit.

Figure 3.49: Schematic profile of a slope failure showing travel distance and travel
distance angle

For this review though, of interest is the research relevant to smaller volume slides more
typical of landslides in soil slopes.
Wong and Ho (1996) analysed the post-failure movement of forty-two landslides in
cut slopes on Lantau Island (Hong Kong) following the severe rainstorm of November
1993. They compared the travel distance beyond the toe of the cut slope, R, to the
height of the slope, Hs, (Figure 3.50) and concluded that:
• For shallow landslides in cut slopes with a relatively flat ground slope at the toe of
the cut, the travel distance (R) ranges from 0.33 to 1 Hs (average of 0.7 Hs).
• If the landslide incorporates a substantial portion of the slope uphill of the cut, R can
increase to up to 1.5 Hs.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.109

• For slopes below the toe of the cut slope greater than about 15 degrees, R exceeding
2 Hs is likely.

Figure 3.50: Travel distance from the toe of the cut, R, versus cut slope height, Hs, for
failures in cut slopes from Lantau Island, Hong Kong (Wong and Ho 1996)

Corominas (1996a), from a database of 204 landslides in soil and rock and covering a
volume range from small (less than 10 m3) to very large, considered the mechanism of
slide initiation and topographical constraints (including confinement of the slide mass
and obstructions to the mass movement) as a means of prediction of the travel distance
angle. His data (all 204 slides) is shown in Figure 3.51 and the results of statistical
analysis are presented in Table 3.22 for a power law fit of the ratio H/L to volume
(Equation 3.11) for various landslide groups. The results indicate there is significant
scatter even within the various landslide groups and sub-groups.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.110

H L = AV B (3.11)

where H and L (in metres) are defined in Figure 3.49, V is the volume of the landslide
mass (in cubic metres), and A and B are constants.
The degree of scatter, particularly for small volume landslides, is emphasised from
the results of Finlay et al (1999) shown in Figure 3.52 for landslides that break-up on
sliding. The range of travel distance angle is very large at the small volume end of the
scale and reflects the significantly greater influence of factors other than volume on
travel distance angle. Fell et al (2000) comment that the large range of travel distance
angle for the smaller volume landslides (from Hong Kong) reflect the different
mechanisms of failure and the degree of obstruction. The very low travel distance
angles for some fills being associated with liquefaction and flow sliding or wash-out
(surface water flow over the fill), and the high values representing cases of dilatant
failure or falls from very steep cut slopes.

Figure 3.51: Landslide volume versus tangent of the travel distance angle for 204
landslides (Corominas 1996a)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.111

Table 3.22: Regression analysis of H/L (Equation 3.11) versus landslide volume
(Corominas 1996a)
Landslide Type Number Path Type A B R2 (*1) Standard
of Slides Error *2
All landslides 204 All 0.897 -0.085 0.625 0.161
Rockfalls 47 All 1.622 -0.109 0.759 0.123
16 Unobstructed 1.469 -0.119 0.924 0.073
Translational 69 All 0.693 -0.068 0.670 0.137
slides 42 Unobstructed 0.719 -0.080 0.796 0.115
Debris flows 71 All 0.972 -0.105 0.763 0.137
19 Confined 0.838 -0.109 0.690 0.171
18 Unobstructed 0.931 -0.102 0.868 0.093
Earthflows and 17 All 0.611 -0.070 0.648 0.131
mudslides 8 Unobstructed 0.603 -0.138 0.908 0.074
1 2
* R is the regression coefficient of the statistical correlation between H/L and slide volume
*2 The standard error applies to log H/L and not H/L.

Figure 3.52: Ratio H/L versus volume for landslides that break-up on sliding (Finlay et
al 1999)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.112

3.9 P OST-F AILURE D EFORMATION – DEVELOPED M ETHODS FOR


ASSESSMENT OF TRAVEL DISTANCE FROM THE DATABASE
In the following sections the results of empirical analyses from the database are
presented, with the purpose of providing empirical correlations for prediction of travel
distance angle that are useful of “rapid” landslides. The analyses concentrate on the
smaller volume (less than 10 6 to 10 7 m3) sub-aerial landslides, as these slides are of
greater concern to engineers and engineering geologists dealing with soil slopes, within
the following material and slope types:
• “Rapid” landslides from failures in dilative soils in cut slopes.
• “Rapid” landslides from failures in dilative soils in natural slopes.
• “Rapid” landslides from failures in contractive soils (flow slides) in loose silty sand
to sandy silt materials.
• “Rapid” landslides from failures in contractive soils (flow slides) in loose dumped
sandy gravels.
• “Rapid” landslides from failures in contractive soils (flow slides) from tailings
dams, in sensitive clays and in hydraulic fill embankment dams.

As will be shown, useful empirical predictive methods require consideration of the


factors influencing the travel, as outlined in Section 3.8.1; in particular the failure
mechanism, material type, mechanism of post-failure movement, degree of confinement
and down-slope geometry. Within each sub-group of “rapid” landslides it is considered
that the dominant factors affecting travel distance vary, and these are discussed in the
following sub-sections.
For larger volume (volumes greater than 10 6 m3) “rapid” landslides in rock slopes
Glastonbury et al (2002) discuss the dominant factors influencing travel distance and
provide guidelines for prediction of travel distance angle (and travel distance).
The statistical analyses undertaken on any set of data initially comprised a test of
significance to determine if the null hypothesis was acceptable (i.e. no correlation
exists) or could be rejected (i.e. accept the correlation exists) using a 95% confidence
interval for the null hypothesis. If the null hypothesis was rejected the goodness of fit
of the correlation can then assessed by the regression coefficient (R2), the higher the
value of R 2 the better the correlation, although the number of data points within the data
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.113

set does have a significant influence on R 2. The following general terms have been
adopted to describe the significance of correlation between factors based on R 2 values:
• R2 < 0.3 Poor correlation
• R2 = 0.3 – 0.5 Minor correlation
• R2 = 0.5 – 0.65 Moderate correlation
• R2 = 0.65 – 0.80 Good correlation
• R2 > 0.80 Excellent correlation

3.9.1 TRAVEL DISTANCE VERSUS SLIDE VOLUME AND INITIAL SLIDE CLASSIFICATION

Figure 3.53 presents the ratio of H/L (tangent travel distance angle) versus volume for
all slides from the database. It includes the approximate limits of the data from Finlay
et al (1999) for small volume landslides in Hong Kong and the mean and 95%
confidence limits for all slides by Corominas (1996a). The slides analysed as part of
this study have been grouped based on the initial landslide type (i.e. flow slides in
contractile soils and the remaining groups as dilative soils) and then sub-grouped based
on factors considered to dominantly influence the travel distance angle. It is evident
from Figure 3.53 that, whilst a general trend of reducing H/L with increase in volume is
evident there is considerable scatter in the data.
Significant departure outside the Corominas (1996a) confidence limits occurs for the
small volume landslides from Hong Kong (Finlay et al (1999) data) and for the strongly
retrogressive landslides (hydraulic fills, sensitive clays and sub-aqueous slopes). Some
sub-groups generally plot below the Corominas mean (flows slides in loose fills from
Hong Kong, flow slides in coking coal stockpiles and confined debris flows) and some
sub-groups above (cut slopes). The confidence limits and data for the smaller slide
volumes in the range most appropriate to engineers and engineering geologists (less
than about 106 m3) cover a broad range of travel distance angle, thereby limiting their
usefulness.
Corominas (1996a) presented data for small and large volume landslides combined,
which had high regression coefficients indicating a good to excellent correlation
between H/L and slide volume. However, further analysis of the Corominas (1996a)
data for debris flows shows (Table 3.23, Figure 3.54 and Figure 3.55) that when the
small volume landslides (less than 106 m3) are analysed separately the regression
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.114

coefficients are significant lower and statistically the correlation is a lot weaker. No
correlation was evident for the confined debris flows.

1.E+01
Finlay et al (1999) data for
cuts, fills and retained slopes
from Hong Kong Corominas (1996a),
95% confidence limits
of data

1.E+00
Corominas (1996a)
data mean
Ratio H/L

1.E-01

1.E-02

1.E-03
1.E+00 1.E+02 1.E+04 1.E+06 1.E+08 1.E+10 1.E+12
Volume (cu.m.)
Cut Slopes (Hong Kong)
Unconfined debris flows in soil and rock (initial failure in dilative soil/rock mass)
Chalk talus (Hutchinson 1988)
Confined debris flows in soil and rock (initial failure in dilative soil/rock mass)
Flow slides in loose fills (Hong Kong)
Flow slides in coal waste spoil piles (South Wales and British Columbia)
Flow slides in coking coal (Hay Point)
Flow slides, chalk cliffs (Hutchinson 1988)
Strongly retrogressive flow slides (sensitive clays, hydraulic fills, tailings dams, sub-aqueous slopes)
Figure 3.53: Ratio H/L versus volume for all slides from database.

Table 3.23: Summary of empirical correlation coefficients (Equation 3.11) for debris
flows by Corominas (1996a)
Travel Volume Number A B R2 Std. *1 Comments
Classification Range of Slides Error
(m3)
All Debris Flows < 1.2x10 10 70 1.005 -0.1056 0.76 0.137 Good correlation
< 10 6 56 1.155 -0.1289 0.37 0.122 Minor correlation
Unconfined < 1.2x10 10 51 1.028 -0.1002 0.82 0.110 Excellent correlation
Debris Flows < 10 6 42 1.158 -0.1219 0.46 0.094 Minor correlation
Confined Debris < 3x109 19 0.866 -1.095 0.69 0.171 Good correlation
Flows < 10 5 14 0.774 -0.0933 0.13 - No correlation
Note: *1 The standard error applies to log H/L and not H/L.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.115

In Figure 3.54 and Figure 3.55 the H/L axis is presented in normal scale to highlight the
broad scatter around the trendline at small volumes compared to the lesser degree of
scatter at larger volumes. It is evident therefore; using correlations of this type for
prediction of the travel distance angle for small volume landslides can lead to
significant errors.

0.8

0.7 Unconfined debris flows


6 3
Vol. Range < 10 m
-0.122
0.6 Y = 1.16 X
2
R = 0.46
0.5
Unconfined debris flows
Ratio H/L

10 3
0.4 Vol. Range < 1.2x10 m
-0.100
Y = 1.03 X
0.3 R2 = 0.82

0.2

0.1

0.0
1.E+00 1.E+02 1.E+04 1.E+06 1.E+08 1.E+10 1.E+12
Volume (cu.m.)
Figure 3.54: H/L ratio versus volume relationships for Corominas (1996a) data on
unconfined debris flows.

0.8

0.7 Confined debris flows


9 3
Vol. Range < 3x10 m
0.6 -0.1095
Y = 0.866 X
2
R = 0.69
0.5
Ratio H/L

0.4

0.3

0.2
Confined debris flows
5 3
Vol. Range < 10 m
0.1 -0.0933 2
Y = 0.774 X , R = 0.13

0.0
1.E+00 1.E+02 1.E+04 1.E+06 1.E+08 1.E+10
Volume (cu.m.)
Figure 3.55: H/L ratio versus volume relationships for Corominas (1996a) data on
confined debris flows.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.116

3.9.2 TRAVEL DISTANCE VERSUS SLIDE VOLUME , DEGREE OF CONFINEMENT OF THE


TRAVEL PATH AND INITIAL SLIDE CLASSIFICATION

3.9.2.1 Travel Distance Beyond the Toe of the Source Area Versus Degree of
Confinement

Prior to presenting the results of the analysis in terms of the ratio H/L, the effect of
degree of confinement on the horizontal distance of travel beyond the toe the source
area is discussed; the purpose being to provide an indication of the actual distance of
travel of the slide mass. For unconfined travel paths the case study data is presented in
Figure 3.56 and for confined and partly confined travel paths in Figure 3.57. Strongly
retrogressive slide types have been excluded.
The findings indicate that the horizontal distance of travel beyond the toe of the
source area increases with increasing slide volume. The distance of travel for confined
travel paths is greater than for unconfined travel paths of similar volume, in the order of
3.5 to 4.5 times for slide volumes in the range 10 3 m3 to 10 6 m3. For most unconfined
travel paths the horizontal distance of travel beyond the toe of the source area is less
than 100 m for slide volumes up to 10,000 m3.
The trendlines in Figure 3.56 and Figure 3.57 are not intended for use in prediction
of the travel distance as they do not consider many of the factors known to influence the
travel of the slide mass. The data serves merely to indicate that, in general, confined
travel paths result in greater distances of travel beyond the source area. The overall
trend with slide volume is to be expected, however, when considering a specific slide
type the correlation is quite poor in most cases. The better correlations are obtained
where the slide is within a particular material type, and the slope of the travel path and
height of the source area is similar, e.g. the flow slides in stockpiled coking coal at Hay
Point.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.117

10000
Cut Slopes (Hong Kong)
. Fill slopes (not flow slides), Hong Kong
Debris flows in natural slopes (Hong Kong)
Horizontal distance of travel beyond the
Flow slides - loose silty sand/sandy silt fill, Hong Kong
Flow slides - Coal mine waste spoil piles
1000
Flow slides - coking coal (Hay Point)
Flow slides - hydraulic fill dams
source area (m)

100

Trendline
0.322
10 y = 4.49 x
2
R = 0.60

1
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Figure 3.56: Travel distance beyond source area versus volume for unconfined travel
paths.

10000
Trendline
.

0.278
y = 27.2 x
Horizontal distance of travel beyond the

2
R = 0.60
1000
source area (m)

100

Confined debris flows in natural slopes, Hong Kong


10
Partly confined debris flows in natural slopes, Hong Kong
Coal mine waste stockpiles (B.C.) - confined, high mobility
Coal mine waste stockpiles - confined, normal mobility
Coal mine waste spoil piles - partly confined
1
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07

Volume (cu.m.)

Figure 3.57: Travel distance beyond source area versus volume for confined and partly
confined travel paths.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.118

3.9.2.2 Travel Distance Versus Degree of Confinement of the Travel Path


Within this section, the data has been analysed to assess the influence of the degree of
confinement of the slide mass in the travel path (in combination with slide volume) on
the travel distance angle. Travel paths have been sub-divided into confined, partly
confined and unconfined. Excluded from this analysis are the strongly retrogressive
landslides (tailings dams, sensitive clays and sub-aqueous slides), hydraulic fill
embankment dams, sturzstroms, chalk talus and Corominas (1996a) debris flows greater
than 10 6 m3 in volume. The flow slides from chalk cliffs and in coking coal stockpiles
have been included but are considered as special groups.
Plots of the data are presented in the format H/L (normal scale) versus slide volume
(log scale). The purpose of presenting the plots in this format is so that the variation
around the trendline is more clearly evident than log-log plots.
The trendlines of travel distance angle versus slide volume and degree of travel path
confinement (Figure 3.58) for “rapid” landslides from failures in dilative soils from
Hong Kong (trendlines from Figure 3.59 to Figure 3.63) indicate that, in general, the
greater the degree of confinement the lower the travel distance angle for a given slide
volume. This finding is in agreement with the data for other landslide specific groups
(e.g. flow slides in coal waste spoil piles from British Columbia) and the research
findings of others (Corominas 1996a; Golder Assoc. 1995; Wong and Ho 1996; Evans
et al 1997).

1.2

1.0 Unconfined travel path


(excl. cut slopes)

0.8
Cut slopes, unconfined
Ratio H/L

travel path
0.6

0.4

Partly confined travel


0.2 path
Confined travel
path
0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Figure 3.58: Comparison of trendlines of H/L versus slide volume differentiated on
travel path confinement for “rapid” landslides from failures in dilative soils, from Hong
Kong.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.119

Figure 3.59 to Figure 3.61 present the H/L ratio versus slide volume for unconfined
travel paths (or travel on open slopes) of “rapid” landslides from failures in soil slopes.
Table 3.24 presents a summary of the empirical analyses for a power law fit between
travel distance angle and slide volume (in the form of Equation 3.11) for various data
groupings based on degree of confinement of the travel path, mechanism of failure,
slope type and material type.
The results indicate that for unconfined “rapid” landslides from sub-aerial slopes
there is a general trend of decreasing travel distance angle with slide volume. However,
within any individual sub-group the slope type, failure mechanism and material type
have a strong influence on the travel distance angle. The correlations between H/L and
volume (Table 3.24) generally range from no correlation to minor. For the larger data
sets the regression coefficients are generally higher, however, there is generally a large
degree of scatter around the trendline.
The flow slides in loose contractive fills have significantly lower travel distance
angles than dilatant failures in fill and also cut slopes (Figure 3.59), even though the
soils are from similar origins (all cases from Hong Kong). The potentially high degree
of brittleness at failure and therefore initially high kinetic energy of the slide mass, and
travel on a low undrained strength liquefied basal layer for these flow slides in fine
grained contractive soils is the likely explanation for the lower travel distance angles.
Flow slides in the coarser soil types (sandy gravels to gravelly sands in coal mine waste
spoil piles) appear to have a travel distance angle to volume correlation more in line
with those of the dilative soil types (Figure 3.59), although there is little overlap in the
volume range of these slide groups to confirm this. Explanation for this observation
could be that for these failures (flow slides in coarser materials) the undrained
brittleness index could be lower than for the finer silty sands (as indicated by the
laboratory results of Dawson (1994) and Eckersley (1986)) and that these materials have
a much higher permeability and therefore have greater potential for dissipation of pore
pressures during flow.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.120

1.2 Cut Slopes (Hong Kong)


Fill slopes (not flow slides), Hong Kong
Debris flows < 1 mil. cu.m. (Corominas (1996a) & Hong Kong)
Flow slides - loose silty sand/sandy silt fill, Hong Kong
1.0
Flow slides - coal mine waste spoil piles

Flow slides
0.8 -0.009
Y = 0.41 X
R2 = 0.02
Ratio H/L

0.6

Dilative soils
0.4 -0.087
Y = 0.956 X
R2 = 0.44

0.2 Flow slides in silty sands, HK


-0.082
Y = 0.67 X
R2 = 0.34
0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)

Figure 3.59: H/L versus volume for “rapid” landslides with unconfined travel paths.

1.2

Cut slopes, HK
-0.067
Y = 0.91 X
1.0
2
R = 0.47

0.8
Unconfined debris
flows (excl. cut slopes)
Ratio H/L

-0.110
Y = 1.06 X
0.6
R2 = 0.41

0.4

0.2

0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Cut Slopes (Hong Kong) Fill slopes (not flow slides), Hong Kong
Debris flows < 1 mil. cu.m. (Corominas 1996a) Debris flows in natural slopes (Hong Kong)

Figure 3.60: H/L ratio versus volume for “rapid” landslides from failures in dilative
soils with unconfined travel paths.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.121

1.2
Loose silty sand to sandy silt fill, Hong Kong Coking coal stockpiles (Hay Point)
Coal mine waste spoil piles (South Wales) Coal mine waste stockpiles (B.C.)
1.0

Loose fills, Hong Kong Coarse grained coal mine


0.8 -0.082
Y = 0.67 X waste spoil piles
2 No correlation
Ratio H/L

R = 0.34

0.6

0.4

Coking coal spoil piles


0.2 -0.187
Y = 1.21 X
2
R = 0.44

0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Figure 3.61: H/L ratio versus volume for “rapid” landslides from failures in contractive
soils (flow slides) with unconfined travel paths.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.122

Table 3.24: Empirical correlation coefficients for power law fit of travel distance angle
to slide volume (Equation 3.11, H L = AV B ) for unconfined “rapid” landslides.
Initial Sub-group Volume No. A B R2 Std. *2 Comments
Landslide Range Error
3
Type (m )
All All (excl. coking < 8x106 161 0.848 -0.072 0.47 0.108 Minor to moderate
Unconfined coal & chalk cliffs) correlation (but
large scatter)
Travel Paths
Silty sand fill and 50 to 8x10 6 38 0.412 -0.009 0.02 - No correlation
Slides in coal mine waste
Contractile Loose silty sand fill 50 to 10,500 16 0.671 -0.082 0.34 0.085 Minor correlation
Soils (Flow (Hong Kong)
Slides) Coal mine waste 3000 to 8x106 22 0.238 0.036 0 - No correlation
Coking Coal 850 to 16,000 9 1.209 -0.187 0.44 0.096 Minor correla tion
All < 136,000 123 0.956 -0.087 0.44 0.090 Minor correlation
All (excl cut slopes) 20 to 136,000 62 1.063 -0.110 0.41 0.093 Minor correlation
Cut slopes (Hong 4 to 52,000 61 0.908 -0.067 0.47 0.081 Minor to moderate
Slides in Kong) correlation

Dilative Fill slopes (Hong 40 to 500 10 0.748 -0.034 0.05 - No correlation


Kong)
Soils*1
Natural slopes 20 to 26000 10 0.877 -0.088 0.27 - No correlation
(Hong Kong)
Corominas (1996a) 36 to 136,000 42 1.158 -0.122 0.46 0.094 Minor to moderate
correlation
Failures in Chalk Cliffs 25,000 to 13 10.98 -0.203 0.76 0.101 Good correlation
(Hutchinson 1988; Golder Assoc. 1.3x106
1995)
Note: No. = number of case studies
*1 Inclusive of defect controlled slides, slides of debris and slides through the soil mass.
*2 The standard error applies to log H/L.

Figure 3.62 and Figure 3.63 present H/L versus slide volume for confined and partly
confined travel paths respectively of “rapid” landslides from failures in soil slopes.
Table 3.25 presents a summary of the empirical analyses for a power law fit between
travel distance angle and slide volume (in the form of Equation 3.11) for the confined
and partly confined travel paths.
The results show a general trend of decreasing travel distance angle with slide
volume for both the confined and partly confined travel paths; however, the correlations
are poor to minor. The no to poor correlation between travel distance angle and slide
volume for individual sub-group (Table 3.25), which indicates that factors other than or
in addition to volume have a significant influence on the travel distance angle for
confined and partly confined travel paths.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.123

1.2
Debris flows - confined (Corominas 1996a)
Coal mine waste spoil piles (B.C.) - normal mobility
1.0 Confined debris flows in natural slopes, Hong Kong
Coal mine waste spoil piles (B.C.) - high mobility
Washouts, Hong Kong
Coal mine waste spoil piles (South Wales)
0.8
Confined travel path
Ratio H/L

-0.060
Y = 0.62 X
0.6 2
R = 0.37

0.4

0.2

0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Figure 3.62: H/L versus volume for “rapid” landslides with confined travel paths.

1.2
Partly confined debris flows in natural slopes, Hong Kong
Coal mine waste spoil piles, partly confined
1.0

Partly confined travel path


-0.039
0.8 Y = 0.60 X
R2 = 0.16 (no correlation)
Ratio H/L

0.6

0.4

0.2

0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Figure 3.63: H/L versus volume for “rapid” landslides with partly confined travel paths.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.124

Table 3.25: Empirical correlation coefficients for power law fit of travel distance angle
to slide volume (Equation 3.11, H L = AV B ) for confined and partly confined “rapid”
landslides.
Degree of Sub-group Volume No. A B R2 Std *1 Comments
Travel Path Range Error
3
Confinement (m )
Partly confined All 40 to 84 0.602 -0.0517 0.27 0.124 Poor correlation
and confined 6x10 6
All 80 to 25 0.600 -0.0386 0.16 0.124 Poor correlation
720,000
Natural slopes (Hong 80 to 11 0.579 -0.0391 0.03 - No correlation
Partly confined
Kong), dilative 3,000
Coal mine waste 20,000 to 14 0.333 0.0085 0 - No correlation
(flow slides) 6x10 6
All 40 to 59 0.617 -0.0602 0.37 0.118 Minor
6x10 6 correlation
Dilative Soils 40 to
- All 13,000 39 0.484 -0.0203 0.02 - No correlation
- Natural slopes, HK 6 0.281 0.0469 0 - No correlation
Confined
- Washouts, HK 19 0.495 -0.0221 0.02 - No correlation
- Corminas (1996a) 14 0.774 -0.0933 0.13 - No correlation
Contractile Soils
- Coal mine waste 40000 to 20 1.071 -0.106 0.29 0.125 Poor to minor
6x10 6 correlation
Note: No. = number of case studies
*1 The standard error applies to log H/L.

From the above analysis of travel distance angle to slide volume and degree of
confinement it is evident that whilst there is general trend of decreasing travel distance
angle with increasing slide volume, other factors such as failure mechanism, slide type,
failure geometry type and material properties have a significant influence on the travel
distance angle.

3.9.3 TRAVEL DISTANCE VERSUS VOLUME , DOWN -SLOPE ANGLE AND SLIDE TYPE
FROM FAILURES IN DILATIVE SOILS

The following sections (and Section 3.9.4) discuss the use of slide volume, down-slope
angle ( α 2 ), slide type and material properties to predict travel distance angle.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.125

3.9.3.1 Cut Slopes


The Hong Kong case studies of “rapid” landslides from failures in cut slopes have been
analysed in detail. The failures includes:
• Slides through the soil mass within residual soils and weathered rock (saprolite).
• Strongly defect-controlled slides within the weathered rock slope (saprolite).
• Slides of debris in colluvial soils from the upper portion of the cut slope.
• Slides through soil mass within fill slopes of typically silty gravelly sands and
clayey silty sands. This group of failures are not flow slides, but considered to be
within initially dilative soils. Only two case studies of this type of failure have been
included.

A significant portion of the empirical analyses on travel distance angle and travel
distance for landslides in Hong Kong has been centred on cut slopes, for which Finlay et
al (1999) and Wong and Ho (1996) have developed useful correlations. The research by
Finlay et al (1999) found that the travel distance, L, was strongly correlated (R2 = 0.85)
to the landslide height, H, and cut slope angle, a cut, (Equation 3.12) for cut slopes where
the slope below the cut is near horizontal. The correlation developed by Finlay et al
(1999) was based on 515 cut slope failures with mean volume of 138 m3 (5th percentile
of 3 m3 and 95th percentile of 446 m3) and mean cut slope angle of 62 degrees (5th
percentile of 45º and 95th percentile of 80º). Given that the power function of H in
Equation 3.12 is very close to 1 the equation can be more simply presented in the form
of H/L (Equation 3.13) without introducing significant error (error of 1 to 5% for cut
slope angles of 30 to 75 degrees and heights of 5 to 50 metres).

L = 1.2853 ∗ H 1.010 * (tan α cut )−0.506 , R2 = 0.85 (3.12)

H / L = 0.78 ∗ (tan α cut )0. 5 (3.13)

The case study data for all cut slopes and those for which detailed slope geometry and
travel distance is available from Hong Kong is presented in Figure 3.64. The subset is
seen to be representative of the overall data set. But, the correlation of travel distance
angle to slide volume alone is minor, as indicated by the significant scatter around the
trendline.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.126

Figure 3.65 presents the same set of data (excluding several case studies where the
run-out was not onto a near horizontal slope at the toe of the cut) with the H/L ratio
normalised against the tangent of the cut slope angle to the power 0.42 (best fit
determined by multiple regression analysis). This shows a moderate to good
correlation, with regression coefficient of 0.62.
This finding is not unexpected when the mechanics of sliding are considered. The
sharp change in slope angle from a steep cut slope angle to a flat slope at the toe is a
significant obstruction to the post-failure travel and would result in a significant loss of
energy of the slide mass. A reduction in the difference in this angle (either by a
flattening of the cut slope angle or steepening of the slope below the cut) would reduce
the loss in energy of the impact at the change in slope and result in increased travel
distance and decreased travel distance angle.
As the volume of the slide mass increases it is considered that the energy lost on
impact per unit volume is reduced. As idealised in Figure 3.66, the initial material
deposited at the toe of the cut reduces the sharp change to a more gradual change in
slope angle, and the debris following would lose less energy in the transition of
movement direction. Finlay et al (1999) did not find that volume was a significant
factor for the travel distance angle from cut slope failures, and it is considered the
reason is the heavy weighting to small volumes in their database, whilst this database
has a wider spread.

1.2
Cut slope failures from Hong Kong
analysed in detail
1.0
Cut slopes failures from Lantau Island
(Wong & Ho, 1996)
0.8
Ratio H/L

0.6

All cut slopes, HK


-0.067
0.4 Y = 0.91 X
2
R = 0.47 Outlier
Detailed analysis
0.2 -0.095
Y = 1.07 X
2
R = 0.43
0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05
Volume (cu.m.)
Figure 3.64: H/L versus slide volume for cut slope cases studies from Hong Kong
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.127

1.0
Type 2 failures in cut slopes

Type 1 failures in cut slopes


0.42

0.8
(H/L) / (tan (cut slope angle))

-0.109
Y = 1.09 X
2
0.6 R = 0.62

0.4

Lower Bound for Type 2 cut


0.2 slopes (refer Figure 3.67b)

0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05
Volume (cu.m.)
Figure 3.65: H/L normalised against tangent of the cut slope angle versus volume for cut
slopes in Hong Kong failing onto a near horizontal slope at the toe.

Figure 3.66: Idealised effect of volume on travel distance for cut slope failures of (a)
small volume compared with (b) significantly larger volumes.

For the cut slope case studies analysed in detail the relationship between travel distance,
R, and cut slope height, Hs, is generally in agreement with the findings from Wong and
Ho (1996). However, the scatter of results is significant and regression analysis
indicates a very weak correlation compared with the correlation of travel distance angle
to cut slope angle and volume shown in Figure 3.65.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.128

Further analysis of the cut slope case studies indicates that other factors have an
influence, but to a lesser degree than volume and cut slope angle, on the travel distance
angle, they are:
• The location of the slide in the cut slope; Type 1, at the top of the cut slope, or Type
2, encompassing virtually the full height of the cut slope (Figure 3.67). In Figure
3.65 the Type 2 cut slope failures plot in the range slightly above to well below the
trendline shown indicating a lower travel distance angle. This finding is to be
expected given that the basal sliding surface of the Type 2 slope category results in a
lower difference in the angle with the slope at the toe of the cut (Figure 3.67b) and
would therefore reduce the loss in energy of the impact.
• Obstructions to the travel of the slide mass, such as buildings. The obstructed slides
result in a loss of energy on impact with the obstruction and therefore it would be
expected that the travel distance angle would be increased due to obstructions. In
Figure 3.65 the obstructed slides generally plot above the trendline, although in
several cases, mainly Type 2 slides they plot below it. Scale effects also require
consideration for obstructions as small volume slides can be obstructed by relatively
small obstructions, which for a larger volume slide are not significant. For example,
the wall in front of a house that obstructed the 85 m3 slide at Chung Shan Terrace
(Slide B) would have had little significance for the 40,000 m3 Po Shan slide of 1972.
• The initial failure mechanism (defect-controlled or slide through soil mass),
geological origin and material type do not show an observable trend in Figure 3.65.
They would therefore appear to have a minor effect on the travel distance angle.
Finlay et al (1999) also found that geological origin and material type had no
discernable influence on the slide mobility.

Figure 3.67: Classification of slope categories (a) Type 1 – failure at top of cut slope,
and (b) Type 2 – failure encompassing the full cut height.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.129

In summary, the dominant factors considered to affect the travel distance angle for
“rapid” landslides from cut slopes failures (from Hong Kong) are the cut slope angle
and volume of the slide mass. The slope angle below the cut also has an influence and
to a lesser extent so does the location of the slide within the cut slope (i.e. Type 1 or
Type 2), and obstructions in the travel path. Flatter cut slope angles, increased volume,
steeper slope angles below the cut and Type 2 cut slopes (encompassing the full height
of the cut) result in a reduced travel distance angle and potentially greater travel
distances. Steeper cut slope angles, smaller volumes, Type 1 cut slopes (above the toe
of the cut) and obstructions in the travel path result in an increased travel distance angle
and potentially reduced travel distances.
Figure 3.65 and Equation 3.14 presents a useful guide to prediction of the travel
distance angle for cut slopes of Type 1 failure geometry (see Figure 3.67) and near
horizontal slopes below the cut, for slide volume in the range 25 to 20,000 m3. For slide
volumes less than 25 m3 the approximation of the corrected Finlay et al (1999)
correlation (Equation 3.13) is recommended. For Type 2 cut slopes the lower bound
line in Figure 3.65 is recommended for predictive purposes.

H / L = 1.09 ∗ (tan α cut )0. 42 ∗ V −0.11 for slide volumes of 25 to 20,000 m3 (3.14)

where α cut is the cut slope angle and V is the estimated slide volume (in cubic metres).
Based on energy principles, where the slope below toe of the cut is not near
horizontal, Figure 3.65 and Equation 3.14 can be used provided that the difference in
slope angle between the cut slope and slope below the toe of the cut is used in Equation
3.14 in place of the cut slope angle (i.e. α cut − α 3 for Type 1 failure geometries and

α cut − α 2 for Type 2 failure geometries, instead of α cut , refer Figure 3.67).

Limitations to the above guidelines include:


• Material type and geological origin. The information presented is considered
suitable for prediction of travel distance angle for colluvium, residual soils and
weathered rocks of granitic and volcanic origin. It is likely to be a reasonable
predictor for slopes in soils of silty sands, clayey silty sands and sandy clays of low
plasticity, and more unreliable for more plastic soils. Extension of its use for cut
slopes in other soils and their parent rocks should be done so with caution, and by
verification against a number of failures in slopes similar to that being considered.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.130

• Slide volume. The case studies on which the analysis is based are limited to
volumes up to 40,000 m3, with only two slides above 10,000 m3. It is considered
that the volume limitation is in the order of 10,000 to 20,000 m3.
• The cut slope angle. From the case studies analysed, the guidelines are considered
appropriate to cut slopes steeper than 33 degrees and flatter than 70 degrees (and
similar differences in slope angle for slopes below the cut that are not near
horizontal).
• Type 2 failures (Figure 3.67b) where the basal angle of the surface of rupture ( α base ,

see Figure 3.67) is more than 15 to 20 degrees steeper than the angle of the slope
below the cut (a 2 for Type 2 failure geometries). The case study identified as an
outlier on Figure 3.64 was a Type 2 cut slope failure with the basal angle of the
surface of rupture almost coincident with the angle of the slope below the cut, and
for these cases there is limited loss in energy due to the slope angle transition.
While there is only one case study data point, the use of the correlations derived
from natural slopes (Section 3.9.3.2) would be more appropriate for prediction of
travel distance angle for these cases.

3.9.3.2 Natural Slopes in Dilative Soils

The analysis of travel distance angle in terms of slide volume (Section 3.9.2) for “rapid”
landslides from natural slopes (in Hong Kong) found a negligible to very weak
correlation, even when the data is sorted based on degree of confinement of the travel
path. Prior to discussing improved methods for prediction of travel distance angle, the
characteristics of the slide mass movement are considered. The characteristics of the
travel and deposition of the slide mass, from the case study data, indicate:
• The distance of travel beyond the toe of the source area for confined travel paths is
significantly greater than for unconfined and partly confined travel paths.
• The smaller volume (less than about 500 m3) “rapid” landslides with unconfined
travel paths generally deposit slide debris along the entire travel path regardless of
the down-slope angle (up to 43 degree slopes) and will terminate on these steep
slopes. Where the travel path extends on to flatter slopes (less than about 10 to 15
degrees) the slide mass will quickly come to rest.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.131

• For the larger volume “rapid” landslides along unconfined travel paths deposition
occurs on slopes as steep as 30 degrees and the slide will flow for significant
distances on slopes less than 10 to 15 degrees before eventually coming to rest.
• “Rapid” landslides along partly confined travel paths generally show similar
characteristics to those on unconfined travel paths. In one case (Lui Pok School,
slide volume 450 m3) some of the characteristics were more like a confined debris
flow with minimal deposition and minor accumulation on slopes of 15 to 30
degrees. The slide mass eventual came to rest after travelling about 40 m on the flat
slopes of the school playground.
• For “rapid” landslides along confined travel paths the length of travel beyond the toe
of the source area can be very long regardless of the slide volume, possibly
reflecting entrainment of water within the flow. In general, these slides accumulated
slide debris on slopes steeper than 15 to 20 degrees, continued to flow on slopes
steeper than about 10 to 15 degrees and travelled for long distances on slopes of less
than 10 degrees. These slides usually terminated either on unconfined relatively flat
slopes in a fan type deposition or were obstructed at the toe of the confined travel
path (e.g. an embankment, reservoir of water or building).

These observations are similar to those of other researchers (Franks 1996; Finlay et
al 1999) who analysed “rapid” landslides from natural terrain from Hong Kong. Franks
(1996) found that for unconfined travel paths deposition starts at an average slope angle
of 27 degrees, with a standard deviation of 6 degrees. For confined debris flows Franks
(1996) comments that deposition will occur on slopes of less than 20 degrees and Finlay
et al (1999) comment that where the confined debris flow accumulates mass (without
deposition) flow will continue on slopes down to 14 degrees.
Analysis of the data on “rapid” landslides from natural slopes found that the down-
slope angle below the source area, a 2, and degree of confinement of the travel path have
a significant influence on the travel distance angle. This is not unexpected in light of
the characteristics of the slide mass movement discussed above. The down-slope angle
below the source area, a 2, is defined in Figure 3.68 and is the average angle of the
portion of the down-slope from below the landslide source area to the point where the
slope begins to flatten out. This definition is somewhat subjective; however, for the
case studies analysed it typically represents at least 50% of the length of travel distance
beyond the toe of the source area (for confined travel paths) and in a number of cases up
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.132

to 100% of this length (for a number of unconfined and partly confined travel paths).
Thus, a 2 incorporates short steeper and flatter sections (e.g., terraces) on the upper
portion of the travel path.

Figure 3.68: Definition of down-slope angle below source area, a 2.

Figure 3.69 to Figure 3.71 present plots of the travel distance angle to down-slope
angle, and trendline of best fit for unconfined, partly confined and confined travel paths
respectively for the Hong Kong case studies. Also shown in these figures is a 1 to 1
trendline between H/L and tangent of the down-slope angle, a 2. No published
information on the down-slope angle was available for the Corominas (1996a) “rapid”
debris flows to incorporate them into the analysis.
The plots (Figure 3.69 to Figure 3.71) indicate a moderate to excellent correlation of
the travel distance angle to down-slope angle, with the travel distance angle increasing
with increasing down-slope angle. The poorer correlation for the partly confined travel
paths is associated with the different types of travel path within the partly confined
group; for example, some slide mass movements are located entirely within relatively
broad gullies and others have short sections of the travel path in a relatively confined
gully.
Comparison of the trendlines based on the degree of confinement of the travel path
(Figure 3.72) shows that they are relatively closely grouped. However, for a specific
down-slope angle below the source area, a 2, the H/L ratio is highest for the unconfined
travel path and lowest for the confined travel path. This relatively small difference in
H/L ratio can translate into a relatively large difference in travel distance for the down-
slope topography represented by the case studies. It is therefore important to consider
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.133

the degree of confinement in travel distance angle predictions. The likely availability of
water in confined travel paths, which can become entrained in the slide mass and result
in greater mobility, is considered the dominant reason for lower frictional losses
apparent for confined travel paths.
The trendlines for the different travel paths are almost coincident with each other
and the 1 to 1 trendline at down-slope angles (a 2) in the range 15 to 22 degrees. These
slope angles are considered to be close to the lower limit of slope angles at which
“rapid” landslides will develop and continue to slide or flow (refer Section 3.5.2.4).
With steeper slopes the kinetic energy of the slide mass is increasing during travel, so
the slide mass has greater kinetic energy to run-out further distances on the lower flatter
slopes below.
The increase in travel distance angle with increasing down-slope angle (a 2) is likely
to be a result of the increasing energy loss in travel of the slide mass. With increasing
down-slope angle the velocity of the slide mass is likely to be greater, resulting in
greater energy loses due to disaggregation of the slide mass.

1.0
Unconfined travel path
0.9
Y = 0.77X + 0.087
2
0.8 R = 0.71
3
Vol. Range = 20 to 26,000 m
0.7

0.6
H/L Ratio

0.5

0.4

0.3

0.2

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Tan (down-slope angle)
Figure 3.69: H/L versus tangent of the down-slope angle, a 2, for unconfined travel
paths, failures in natural slopes from Hong Kong.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.134

1.0
Partly confined travel path
0.9
Y = 0.691X + 0.0862
2
0.8 R = 0.52
3
Vol. Range = 80 to 3,000 m
0.7

0.6
H/L Ratio

0.5

0.4

0.3

0.2

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Tan (down-slope angle)
Figure 3.70: H/L versus tangent of the down-slope angle, a 2, for partly confined travel
paths, failures in natural slopes from Hong Kong.

1.0

0.9
Confined travel path
0.8 Y = 0.54X + 0.147
2
0.7
R = 0.85
3
Vol. Range = 50 to 13,000 m
0.6
H/L Ratio

0.5

0.4

0.3

0.2

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Tan (down-slope angle)
Figure 3.71: H/L versus tangent of the down-slope angle, a 2, for confined travel paths,
failures in natural slopes from Hong Kong.

The scatter around the trendlines in Figure 3.69 to Figure 3.71 is not necessarily
associated with slide volume. Only for the unconfined travel path events did the larger
volume slides plot below the trendline; for partly confined and confined travel paths no
trend with slide volume was evident. No trend was evident for major obstructions to the
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.135

slide mass either. The reasons for the scatter are considered to be associated with
relatively small obstructions to the slide mass particularly for the smaller volume slides,
the flow behaviour, the potential availability of water that can become entrained in the
slide mass and differences in material properties (geological origin had no influence).
Multiple regression analyses of H/L with slide volume and down-slope angle below
the source area, a 2, were also carried out; however, no correlation was found for the
partly confined and confined travel paths, and only a poor correlation was obtained for
the unconfined travel path.

1.0
All travel paths Unconfined travel paths
0.9 Y = 0.74X + 0.078
2
0.8 R = 0.66
3
Vol. Range = 20 to 26,000 m
0.7

0.6
H/L Ratio

0.5

0.4 Partly confined


travel paths
0.3

0.2 Confined travel paths

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Tan (down-slope angle)
Figure 3.72: Natural slopes (Hong Kong). Trendlines for H/L versus tangent of the
down-slope angle, a 2, for all types of travel path.

From the study it is apparent that the down-slope angle below the source area, a 2,
presents a useful method for prediction of the travel distance angle for “rapid”
landslides from natural slopes. Equations 3.15, 3.16 and 3.17 represent the trendlines
for unconfined, partly confined and confined travel paths respectively from Figure 3.69
to Figure 3.71.

Unconfined H / L = 0.77 ∗ (tan α 2 ) + 0.087 Std. Error = 0.095 (3.15)

Partly confined H / L = 0.69 ∗ (tan α 2 ) + 0.086 Std. Error = 0.110 (3.16)


Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.136

Confined H / L = 0.54 ∗ (tan α 2 ) + 0.147 Std. Error = 0.027 (3.17)

where a 2 is the down-slope angle below the source area (Figure 3.68).
These equations are also applicable to Type 2 cut slope failures (Figure 3.67b)
where the basal angle of the surface of rupture is less than about 10 degrees (and
possibly up to 15 to 20 degrees) steeper than the angle of the slope below the cut. The
analysis for unconfined travel paths incorporated two case studies of cut slopes that
complied with this criterion.
For prediction of the travel distance angle the following steps are recommended:
• Approximate estimation of the travel distance angle based on the statistical mean of
the travel distance angle from slides in similar soil types with similar degree of
confinement of the travel path (refer Table 3.29).
• Transposing of the travel distance angle onto a long section of the potential slide and
travel path.
• Estimation of the down-slope angle below the landslide source area, a 2, from this
section. The length over which the estimate is made should represents at least 50%
of the length of travel beyond the toe of the source area.
• Prediction of the travel distance angle based on assessment of the degree of
confinement of the travel path from either Equations 3.15 to 3.17 or Figure 3.69 to
Figure 3.71.

Limitations to the above guidelines are:


• Slide volume. The case studies on which the analyses are based is limited to
volumes up to 26,000 m3 and a limitation of 25,000 m3 is considered appropriate at
this stage until further verification is undertaken for larger volume slides.
• Material type and geological origin. The guidelines are considered appropriate to
similar soil types to that analysed, i.e., colluvial and residual soils derived
predominantly from granitic and volcanic (rhyolite and fine to coarse ash tuffs)
rocks. The colluvial soil types cover a broad range of soil types from dominantly
finer grained silty sands to dominantly gravel to boulder size, both with low clay
contents (less than 10%). The number of case studies within residual soils is
limited, and the material types were typically silty sands to clayey silty sands to
sandy silts). The method should be applicable to similar soils but of different
geological origin, but not to more plastic or high clay content soils.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.137

3.9.4 TRAVEL DISTANCE VERSUS VOLUME , DOWN -SLOPE ANGLE AND SLIDE TYPE
FROM FAILURES IN CONTRACTILE SOILS

3.9.4.1 Fills Constructed of Loose Silty Sands


The analysis of flow slides in loose fill slopes of silty sands to sandy silts with low clay
content (from Hong Kong), Section 3.9.2, indicated that a minor correlation exists
between travel distance angle and slide volume (Figure 3.73). Further analysis of the
smaller data set, for which detailed information was available, indicated some
correlation exists between the travel distance angle and down-slope angle below the
source area, a 2. However, given the relatively low regression coefficient (R2 of 0.50)
for the limited number of case studies (7) the correlation was statistically unacceptable.

0.7
Flow slides in loose silty sand fill
-0.082
0.6 Y = 0.67 X
2
R = 0.34
Vol Range = 50 to 10,000 cu.m.
0.5
H/L Ratio

0.4

0.3

0.2 Flow slides in loose silty sand fill (Hong Kong) -


detailed analysis

0.1 Additional flow slides in loose fill from Hong Kong


(Wong & Ho, 1996; Sun, 1998)

0.0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05
Volume (cu.m.)
Figure 3.73: H/L versus slide volume for flow slides in loose silty sand to sandy silt
fills.

Factors other than slide volume influencing the travel distance angle are the relative
density of the fill, obstructions along the travel path and presence of water. The effect
of these factors on the potential travel of the slide mass are:
• Relative density. The lower the relative density of the fill, in comparison to similar
fill types, the lower will be the residual undrained strength on liquefaction and
higher the undrained brittleness index. Therefore, at failure the slide mass will have
a higher kinetic energy than for comparative fills of higher relative density that are
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.138

still susceptible to liquefaction. Thus, the lower the relative density the lower the
expected travel distance angle.
• Obstructions to the travel of the slide mass will reduce its kinetic energy and
therefore reduce the travel distance and increase the travel distance angle.
• Free water, if available, can become entrained in the slide mass and for low
cohesion materials will reduce the strength of the slide mass and potentially lead to
increased travel distance and decreased travel distance angle.

For prediction of the travel distance angle of flow slides in loose silty sand to sandy
silt fills with low clay content (less than 10%) two methods should be used:
(a) For slide volumes between 50 m3 and 500 m3 use the mean and standard deviation
from Table 3.29;
(b) For slide volumes between 500 m3 and 10000 m3 use Figure 3.73 and Equation
3.18.

In each case an appropriate degree of conservatism should be used.

H / L = 0.67 ∗ V −0.082 (standard error of log H/L = 0.085) (3.18)

Limitations to the above guidelines are:


• Slide volume. The case studies on which the analyses are based are limited to
volumes up to 10,000 m3.
• Material type and geological origin. The guidelines are considered appropriate to
similar soil types to that analysed; i.e., silty sands to sandy silts with fines contents
of 10 to 50% and clay contents less than 10% derived from completely to highly
decomposed granitic and volcanic rocks. They should apply other silty sandy soils,
e.g. those from alluvium, colluvium or derived from sandstone.
• Density and/or placement method. The compaction limit for static liquefaction in
these soils is 85% of Standard Maximum Dry Density (HKIE 1998), and the fills
were generally placed by end dumping without any formal compaction.

3.9.4.2 Coarse-grained Mine Waste Spoil Piles on Hillsides


Golder Assoc. (1995) analysed the post failure travel of 40 “rapid” flow slides from coal
mine waste spoil piles in British Columbia, for which the scatter of the travel distance
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.139

angle was large with respect to slide volume (Figure 3.74 ). They derived the
relationship between the excess travel distance, Le, and slide volume (Equation 3.10)
that incorporates a spreading ratio, R (where R = Le / B , B is the breadth of the deposit),
based on the actual case studies and taking into account the degree of confinement of
the travel path. Their analysis was based on the assumption of a correlation between
depositional area and volume (Equation 3.9) and the assumption that Le approximated
the length of the deposit.
Golder Assoc. (1995) identified a group of highly mobile events where the travel
path was confined and the slide debris over-rode suspected liquefiable organic materials
mantling the lower slopes and gullies of the confining channels. This group of slides
was separate from the “normal” mobility events. Golder Assoc. (1995) derived
expressions for Le in terms of slide volume for each group (Table 3.26) and found a
standard deviation of 30% between the actual Le and calculated Le (excess travel
distance) from the correlations given in Table 3.26.
As part of this study, further analysis of the Golder Assoc. (1995) case studies was
undertaken to convert their predictions of the excess travel distance, Le, into predictions
of the H/L ratio for comparison against analyses using other methods. The percentage
differences of the mean and standard deviation for the H/L ratio between the actual and
predicted ratios and the standard error are given in Table 3.26.

Table 3.26: Summary of empirical correlations for flow slides in coal waste spoil piles
in British Columbia
Slide Equations of Best Fit Difference Between Actual and
Mobility (Golder Assoc. 1995) Predicted H/L Ratio
Mean Std. Dev. Standard Error of
H/L prediction
High mobility A = 22 .8 ∗ V 0.662 5% 11% 0.015

(
Le = 22.8 ∗ R ∗ V 0.662 )
0 .5

Normal A = 74 .5 ∗ V 0.5 -2% 12% 0.042


mobility
(
Le = 74.5 ∗ R ∗ V 0.5 )
0 .5

Note: Std. Dev. = standard deviation

The difficulty with analyses of this type is that in order to predict the travel distance
angle of a potential landslide, estimation of the spreading ratio is required. For the flow
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.140

slides in coal waste spoil piles from British Columbia analysed by Golder Assoc. (1995)
the variation in spreading ratio ranged from 0.5 to 100, although there is much less
variability within the morphological sub-groups. The inaccuracy associated with
estimation of the spreading ratio for a potential landslide detracts from using such
correlations and other methods (discussed below) are preferable.
With a view to developing improved methods for prediction of travel distance
further analyses were undertaken of flow slides in coal mine spoil piles incorporating
the case studies from British Columbia and South Wales. It is considered that given the
similarity of the particle size distribution of the liquefaction susceptible materials, the
loose nature resulting from the two methods of placement and that both are on hill
slopes, that methods of prediction of travel distance angle should be applicable to both
sets of case studies. The coking coal stockpile case studies from Hay Point were
excluded due to the low specific gravity (SG approximately 1.35 t/m3) of coking coal
even though their particle size distribution was similar to that of the coal mine spoil pile
case studies.
The long section profile of the travel paths for the flow slides in South Wales was
generally different to that for the flow slides in British Columbia. At South Wales the
hill slope angles (8 to 27 degrees for the case studies) generally maintained a relatively
constant slope, only flattening off toward the toe. In a number of cases (slightly less
than half) the travel of the flow slide terminated on the hill slope. For the remaining
cases the flow reached the toe of the hill slope, but did not generally travel a great
distance beyond the toe, it either came to rest on the flatter slopes or else was obstructed
by embankments or channels in the toe region.
In contrast, the topography in the region of the British Columbian coal mines
consists of glacially over-steepened valleys with typically very steep upper slopes (up to
35 to 50 degrees) that decrease gradually to relatively flat slopes in the toe region of
these valleys, giving the down-slope a markedly curved shape. The travel path for these
slides varied dependent on the position of the toe of the spoil pile in relation to the toe
of the valley, varying from initially relatively steep in some cases to relatively flat in
other cases. In a significant number of cases the flow slide travelled for long distances
(more than 1000 m in several cases) on relatively flat slopes of 0 to 10 degrees. The
high kinetic energy of the slide mass as a result of the steep topography and large height
of the spoil piles (100 to 400 vertical metres) and the potentially liquefiable foundation
soils are dominant factors for the long travel distances on relatively flat slopes for
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.141

several of these slides. In contrast, the vertical height of the spoil piles for the South
Wales case studies typically ranged from 30 to 70 m.
The case study data is presented in Figure 3.74 in the form of H/L versus slide
volume. The two case studies with H/L ratios exceeding 0.6 were considered as outliers
to the group and therefore omitted from the analysis. Fforchman was also omitted as it
was considered not to have initiated as a flow slide. Whilst it does not appear as an
outlier on Figure 3.74 it does so when plotted in the form of H/L versus down-slope
angle below the toe of the spoil pile. The earlier analysis based on degree of
confinement of the travel path (Section 3.9.2) indicated that a poor correlation existed
between the travel distance angle and volume for confined travel paths and no
correlation (with volume) for unconfined and partly confined travel paths.
Further analyses were undertaken of H/L ratio versus the down-slope angle below
the toe of the spoil pile, a 2. The analysis showed (Table 3.27 and Figure 3.75) a general
trend of decreasing H/L ratio (and travel distance angle) with decreasing down-slope
angle below the toe of the spoil pile (a 2) where the angle, a 2, is taken as the average
slope from the toe of the spoil pile for a distance of at least 30 to 40% of the total
distance travelled beyond the toe. The regression coefficients indicate minor
correlations with down-slope angle for confined and unconfined travel paths. No
correlation was statistically evident for the partly confined travel paths. Comparing the
correlations for confined travel paths, the correlation based on down-slope angle is
stronger than for slide volume.
Multiple regression analyses of the tangent of the travel distance angle were
undertaken with respect to a combination of slide volume, down-slope angle below the
toe of the spoil pile and height of the spoil pile. However, no correlations were evident.
It was thought that the height of the spoil pile might have had some discernable effect
on the travel distance angle due to kinematic effects.
The case studies were also analysed by assuming no correlation with volume or
down-slope angle. Table 3.29 (in Section 3.9.5) presents the mean and standard
deviation of the data, and the estimated confidence interval of the standard deviation of
the population. The findings indicate, as shown in Figure 3.74, that the confined, high
mobility flow slides have a lower mean H/L ratio than the remainder of the cases. For
the cases other than confined, high mobility (i.e. partly confined and unconfined travel
path, and normal mobility confined travel path) the mean and standard deviation were
similar.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.142

0.9
British Columbia - unconfined South Wales - unconfined
0.8 British Columbia - partly confined South Wales - partly confined
British Columbia - confined, normal mobility South Wales - confined
0.7 British Columbia - confined, high mobility

Outliers
0.6
Fforchman
Ratio H/L

0.5

0.4

0.3

0.2

0.1

0.0
1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)
Figure 3.74: H/L versus slide volume for flow slides in coarse-grained coal mine waste
spoil piles.

0.7
Confined - high mobility
Confined - normal mobility
0.6
Partly confined - normal mobility
Confined to unconfined travel paths Unconfined - normal mobility
Y = 0.54 X + 0.16
0.5
2
R = 0.37
7 3
Volume range = 3000 to 10 m
Ratio H/L

0.4

0.3

0.2

Confined travel path (normal and high mobility)


0.1 2
Y = 0.57 X + 0.13, R = 0.49
7 3
Volume range = 3000 to 10 m
0.0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
Tan (down-slope angle)

Figure 3.75: H/L versus the tangent of the down-slope angle below the toe of the spoil
pile, a 2; flow slides in coarse-grained coal mine waste spoil piles.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.143

Table 3.27: Statistical summary of H/L to tan a 2 correlation for flow slides in coarse-
grained coal mine waste spoil piles.
Landslide Correlation No. R2 Std. z *1 Comment
Group / Cases Error
Sub-group
All cases H L = 0.53 tan α 2 + 0.13 35 0.37 0.068 4.0 Minor correlation

Confined travel H L = 0.57 tan α + 0.13 17 0.49 0.065 3.2 Minor correlation
2
path - all
Confined travel H L = 0.35 tan α + 0.14 8 0.53 0.027 2.1 Moderate correlation.
2
path – high The z test indicates
mobility the correlation is very
weak.
Unconfined H L = 0.59 tan α 2 + 0.17 11 0.39 0.058 2.1 Minor correlation
travel path
Note: Std. Error = Standard error R2 = regression coefficient
*1 z is the statistic for test of null hypothesis (accept correlation for z > 1.96 or z < -1.96)

The factors affecting the travel distance angle for flow slides in loose, coarse grained
coal mine waste spoil piles are:
• The materials mantling the down-slope travel path. Saturated or near-saturated soils
susceptible to liquefaction on rapid loading can significantly reduce the travel
distance angle. For the flow slides in British Columbia liquefaction susceptible
materials were suspected for the high mobility (> 1 km distance of travel beyond the
spoil pile toe) of a number of flow slides in confined valleys.
• Longitudinal down-slope section. The steeper the initial down-slope angle generally
the larger the travel distance angle.
• Degree of confinement, height of the spoil pile and slide volume are considered to
affect the travel distance angle, although their influence appears to be over-
shadowed by other factors.
• Relative density and undrained brittleness index of the loose dumped waste material
are considered to have an affect on the travel distance angle, but cannot be
quantified.

For prediction of the travel distance angle of flow slides in coarse grained dumped
mine waste deposits on hillsides it is recommended that:
• For longitudinally curved down-slopes, where estimation of the angle a 2 can be
difficult, use the mean and standard deviations from Table 3.29. Take into
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.144

consideration the potential for liquefaction of the material mantling the travel path
where the travel path is likely to be confined. Otherwise use the values for normal
mobility events disregarding the degree of confinement of the travel path.
• Where the down-slope angle is relatively consistent below the toe of the spoil pile
use the down-slope angle correlations given in Figure 3.75.

Limitations to the above guidelines are:


• Slide volume. The case studies on which the analyses are based cover a broad range
of volume from 3000 to 8 x 10 6 m3.
• Material type. The guidelines are considered appropriate to similar material types to
that analysed; i.e., spoil piles of sedimentary rock types composed entirely or partly
of sandy gravels with low silty fines content (less than 3 to 5%) dominantly from
argillaceous rock types. It should apply to materials of similar grading derived from
other rock types.
• Material density/placement method. Density is a very significant factor in
susceptibility of the material to static liquefaction. The guidelines are considered
appropriate to spoil piles formed by end dumping from low height at the crest or
peak of the spoil pile. Spoil piles formed by dumping from height, such as by
dragline, are not likely to be susceptible to flow slide.
• Hillside topography. The formation of the spoil pile on hillsides is a feature of all
case studies. Further details are given in Section 3.4.2.

3.9.4.3 Retrogressive “Rapid” Landslides


The soil types or sub-groups within which slide retrogression is a dominant mechanism
include sensitive clays, tailings dams, submarine slopes and very loose sub-aqueous fill
deposits. The significant feature for retrogression of these slides is continued back-
scarp instability as discussed for sensitive clays in Section 3.4.3.
As shown in Figure 3.53 (and Figure 3.76 for retrogressive slides only) the travel
distance angle for these slide groups is very small compared with “rapid” landslides in
other soils. The dominant factors contributing to this are considered to be:
• The retrogressive nature of the slide. Retrogression distances of several hundred
metres are typical for sensitive clays (up to 6 km has been recorded, Mitchell and
Markell (1974)) and tailings dams. The distance of retrogression, in a number of
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.145

cases, represents a significant percentage of the total slide length, L (15 to 70% for
tailings dams, 10 to 50% for sensitive clays). Therefore, its influence on reducing
the calculated travel distance angle can be significant.
• The significant loss of strength on liquefaction and remoulding (particularly for
sensitive clays) and the high volume of water entrained in the soil structure. These
factors contribute to the fluid like characteristics of the remoulded soil and
potentially large travel distances. In addition, the greater the flow potential of the
remoulded soil the greater the likelihood of continued back-scarp instability due to
exposure of the full back-scarp height (Leroueil et al 1996).
• The relatively low permeability of tailings and quick clays and therefore the likely
low rate of dissipation of pore pressures during flow.
• The post-failure flow behaviour. For the sub-aerial slope failures (tailings and
sensitive clays) the liquefied material has fluid like characteristics and the flow
behaviour is reported (Jeyapalan et al 1983a; Blight 1997) to be laminar on
relatively flat slopes and turbulent for the more “rapid” flows on steeper slopes. For
laminar flow conditions a Bingham plastic fluid model is reported to give a
reasonable approximation of the flow (Jeyapalan et al 1983a; Blight 1997).
• Entrainment of free water into the slide mass will significantly alter the flow
properties of the fluid. For sub-aerial laminar type flows entrainment of water into
the lower portion of the flow can significantly reduce the yield strength at the base
of sliding resulting in much greater travel distances. In several cases confinement of
the flow into running streams and rivers resulted in development of hyper-
concentrated stream flows that travelled for distances in excess of 20 km on
relatively flat slopes; the failures at Bafokeng, South Africa in 1974 and Stava, Italy
in 1985 are examples.

Several case studies of “rapid” flow slides from tailings dams and sensitive clays
have been particularly destructive, resulting in significant loss of life and extensive
damage to property. These two groups of retrogressive landslides therefore deserve
further discussion on the factors influencing their post-failure travel behaviour.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.146

0.35
Tailings Dams
Hydraulic fills (sub-aqueous)
0.30
Submarine Slopes (Edgars & Karlsrud 1983)
Sensitive Clays
0.25
Hydraulic fill embankment dams
Ratio H/L

0.20

0.15

0.10

0.05

0.00
1.E+00 1.E+02 1.E+04 1.E+06 1.E+08 1.E+10 1.E+12
Volume (cu.m.)
Figure 3.76: H/L versus volume for retrogressive “rapid” slides

(a) Flow Slides from Tailings Dams


For tailings dams the risk to life and property is down-slope of the failure. The
properties and factors of those flow slides from tailings dams that were particularly
destructive or had the potential to be particularly destructive are:
• The amount of water stored within the impoundment at the time of failure. Large
volumes of water were reportedly stored within the impoundments of Stava,
Bafokeng, Merriespruit and Buffalo Creek at the time of failure. This free water is
available for entrainment into the flowing tailings and will significantly influence
the flow properties leading to greater travel distances.
• The state of operation. All case studies of statically triggered “rapid” flow slides
from tailings dams were active at the time of failure. Whilst this is not a significant
factor leading to destructive (or potentially destructive) landslides in tailings dams
triggered by static liquefaction, freshly deposited tailings do have a greater potential
for flow. The state of operation is also a significant factor for earthquake triggered
landslides in tailings dams (Tronsoco 1988, 2000).
• Presence of water or wet foundation on the travel path (as discussed above).
• Confinement or partial confinement of the flow. This is particularly significant
where the tailings enter an actively flowing watercourse. For Stava, Buffalo Creek
and Bafokeng (all with travel distances in excess of 20 km) it is considered that the
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.147

flow transformed into a turbulent stream-flow as a result confinement into a flowing


watercourse.
• Down-slope angle. For most of the case studies analysed the gradient of the down-
slope was less than 1.5 degrees. In the case of Stava the down-slope angle averaged
4.5 to 7.6 degrees and the velocity of the flowing mass increased from an average of
8 m/sec over the first 4 km up to 12 to 25 m/sec for the subsequent 20 plus
kilometres. For earthquake-triggered failures it is suspected (from the available
information) that the large travel distances recorded (5 to 12 km in a number of
cases) were associated with down-slope angles that were as shallow as 3 to 5
degrees.

Travel distance angle predictions based on empirical correlations are not


recommended for flow slides in tailings dams. Predictions based on numerical
modelling are considered more appropriate (Section 3.10.3).

(b) Flow Slides from Landslides in Sensitive Clays


For retrogressive landslides in sensitive clays the greatest destructive potential is to
persons and property up-slope within the area encompassed by the retrogression of the
landslide. Leroueil et al (1996) comment that the distance of retrogression is very
difficult to predict, but that it has a tendency to increase with increasing liquidity index.
Figure 3.77 shows a broad degree of variation in the distance of retrogression relative to
the liquidity index for the case study data from Mitchell and Markell (1974) and other
published case studies included as part of this study.
Mitchell and Markell (1974) indicate some correlation of retrogressive distance with
the stability number, Ns (Ns = ?H/cu, where H = slide depth, cu = undrained shear
strength and ? = bulk unit weight), as shown in Figure 3.78. Trak and Lacasse (1996),
for eight flow slides in sensitive clays, showed some correlation between the ratio of
retrogressive distance to slide depth and stability number (slightly different to the
stability number used by Mitchell and Markell). However, addition of a number of flow
slides in sensitive clays from Mitchell and Markell (1974) and from the published
literature show this to provide a poor correlation (Figure 3.79).
In conclusion, the existing empirical based methods for prediction of the
retrogression distance of “rapid” flow slides in sensitive clays, based on liquidity index
and stability number, are relatively inaccurate and should be used cautiously. No
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.148

improvement to the current empirically based methods is suggested from analysis of the
case study data.

4.5
Flow slides, Ip < or = 20 (data from this study)
4.0 Flow slides, Ip > 20 (data from this study)
Flow slide, Ip < or = 20 (Mitchell & Markell 1974)
3.5 Flow slide, Ip > 20 (Mitchell & Markell 1974)
Liquidity Index, IL (%) .

Flow slide, Ip not known (Mitchell & Markell 1974)


3.0 Slide with limited flow or single slide (Mitchell & Markell 1974)

2.5

2.0

1.5

1.0

0.5 Note: Ip = plasticity index

0.0
0 500 1000 1500 2000 2500
Retrogression Distance, Re (m)
Figure 3.77: Distance of retrogression versus liquidity index for slides in sensitive clays

Figure 3.78: Distance of retrogression based on stability number for landslides in


sensitive clays (Mitchell and Markell 1974)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.149

30

25
.
Stability Number, N s

20
Correlation suggested by Trak and
Lacasse (1996) for flow slides
15

Flow slides, Ip < or = 20 (data from this study)


10 Flow slides, Ip > 20 (data from this study)
Flow slide, Ip < or = 20 (Mitchell & Markell 1974)
Flow slide, Ip > 20 (Mitchell & Markell 1974)
Flow slide, Ip not known (Mitchell & Markell 1974)
5
Slide with limited flow or single slide (Mitchell & Markell 1974)
Note: Ip = plasticity index
0
0 25 50 75 100 125 150
Retrogression Distance / Slide Depth (Re/D)

Figure 3.79: Ratio of retrogression distance to slide depth versus stability number

3.9.4.4 Hydraulic Fill Embankment Dams

The post failure analysis of the “rapid” flow slides from failures in hydraulic fill
embankment dams included the failure of Wachusett dam in loose dumped sands due to
similarities in material type, the failure mechanism (flow liquefaction) and the post-
failure deformation behaviour. A summary of the slide properties for the 5 case studies
is given in Table 3.28.

Table 3.28: Summary of slide properties of flow slides in hydraulic fill embankments
Name *1 Material Type Embankment Slide TDA
(Liquefiable Zone) Slide Height Slope Volume (º) *2
Location
(m) (º) (m3)
(shoulder)
Sheffield Dam Silty sands to sandy silts downstream 7.6 22 16 x 103 6.8
Lower San Silty sands to sandy silts upstream 35 21.8 400 x 10 3 9.9
Fernando Dam
Calaveras Dam Not clear what part of upstream 56 18 600 x 10 3 7.3
embankment liquefied,
possibly the siltier and sandier
fractions
Fort Peck Dam Fine sand upstream 58 12.6 8000 x 103 3.8
Wachusett Dam * 3 Fine sand to silty sand upstream 24.4 22 47 x 103 7.9
Note: *1 For all slides the travel path was unconfined and on near horizontal slopes
*2 TDA = travel distance angle, a
*3 Constructed of loose dumped sand
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.150

The limited number of case studies makes empirical methods of analysis based on
historic events difficult to quantify statistically. Whilst regression coefficients may be
relatively high, statistically the correlation may not be acceptable, or only just
acceptable.
The distance of travel beyond the toe of the source area shows an excellent
correlation with slide volume (Figure 3.56) and also with embankment height. The
goodness of fit of these correlations is probably somewhat fortuitous given the small
data set, however, in all cases the failure geometry is similar (all are Type 2 geometry)
and the travel path is unconfined and on near horizontal slopes, and therefore some
correlation to embankment height and slide volume would be expected. The slide
volume has a positive correlation with embankment height (Table 3.28).
When the data is plotted in the form H/L to slide volume the data points plot at and
below the lower 95% confidence interval of Corominas (1996a). This is generally
lower than for most sub-aerial “rapid” landslide groups except for the retrogressive flow
slides in sensitive clays and tailings dams. Figure 3.76 presents the case studies plotted
with the retrogressive flow slides and shows the low H/L ratio of these events (0.07 to
0.18) and the lack of correlation of travel distance angle to slide volume.
Due to the significance of these structures and the potential catastrophic
consequences in the event of an embankment breach, numerical analysis of the post-
failure travel is recommended. Laboratory testing of high quality samples should be
undertaken to assess the potential for liquefaction and provide estimates of residual
undrained strengths. Earthquake is considered to be the most likely trigger of a
potential flow slide event in hydraulic fill embankment dams given that these
construction methods are no longer used in dam construction. However, the guidelines
for post failure analysis are considered applicable to embankments on liquefaction
susceptible foundations and also other hydraulic fill structures.

3.9.5 SUMMARY OF METHODS FOR PREDICTING TRAVEL DISTANCES

Previous researchers have combined the smaller volume slides typical of soil slopes
with the significantly larger volume slides observed in rock slopes and derived
correlations based on slide volume for prediction of the travel distance angle. It has
been found that correlations of this type give poor predictions for the smaller volume
slides.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.151

The analysis of travel distance angle with volume, failure mechanism and degree of
confinement of the travel path found relatively weak statistical correlations (Section
3.9.2). Improvement in the accuracy of predictive methods was obtained when the data
was divided further to take into consideration slope type (cut, fill or natural slope),
material type, initial failure mechanism, down-slope angle below the source area, α 2 ,
the cut slope angle (for cut slopes), travel path confinement or slide volume, or a
combination of these. The main sub-groupings of “rapid” landslides considered were
from:
• Cut slopes (Section 3.9.3.1) in saprolitic, residual soils and some colluvial soils
consisting typically of gravel to boulder sized weathered rock fragments in a silty
sand matrix, with low clay content.
• Natural slopes (Section 3.9.3.2) mainly in cohesionless to low cohesion colluvial
soils. These cases studies were further separated based on degree of confinement of
the travel path.
• Flow slides in fills constructed of loose silty sand with low clay content (Section
3.9.4.1).
• Flow slides in coarse-grained (sandy gravels) loose dumped mine waste fills
(Section 3.9.4.2). These case studies were of the coal mine waste spoil pile failures
in British Columbia and South Wales, and were further divided based on degree of
confinement of the travel path and materials mantling the down-slope travel path.

Guidelines on predictive methods for the travel distance angle for each of the above
sub-groups are given in the referenced chapter sections and are summarised in Table
3.30. Table 3.30 briefly summarises the method of prediction and references the
relevant figures, tables and equations where the correlations can be found, and the
volume range over which the correlation is considered applicable. Limitations on the
application of the correlation are discussed at the end of the chapter sub-sections
referenced above.
For a number of sub-groups of landslides analysed in Sections 3.9.3 and 3.9.4 no or
poor correlations were obtained for the travel distance angle with respect to either slide
volume or the down-slope angle, a 2. For these slide groups it has been recommended to
use the mean and standard deviation of the H/L data specific to that group for prediction
of the travel distance angle. Table 3.29 presents the results of statistical analysis of
these slide groups and includes ranges for slide volume and H/L ratio of the case studies
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.152

analysed. Also included are the data for “rapid” landslides from natural slopes for
preliminary predictive estimates of travel distance angle.
Although not discussed separately in Section 3.9, statistical results from analysis of
the slide groups flow slides in coking coal, dilative failures in fills constructed of silty
sands with low clay content and washout failures of silty sand fills are included in Table
3.29. Further data on the post failure travel of “rapid” landslides in these sub-groups are
given in Table 3.24, Table 3.25, and Figure 3.59 to Figure 3.63. The mean and standard
deviations given in Table 3.29 for these slide groups are a basis for prediction of the
travel distance angle.
Retrogressive flow slides in tailings dams and sensitive clays (Section 3.9.4.3), and
flow slides in hydraulic fill embankment dams (Section 3.9.4.4) have been considered
separately. For flow slides in tailings dams and hydraulic fill embankment dams
empirical methods of predictions are considered not appropriate and numerical methods
of analysis are recommended.

3.10 P OST- FAILURE NUMERICAL M ODELLING - M ETHODS AND R ESULTS


The following sub-sections summarise the results of numerical modelling applied to the
post failure travel of groups of “rapid” landslides from Hong Kong, coal waste spoil
piles in British Columbia and tailings dams. Fell et al (2000) present a summary on the
background and available methods for numerical modelling of the post-failure
behaviour of “rapid” slides. The method that shows the most promise at this time is the
DAN model (Hungr 1995).
The DAN model is a universal continuum model based on the Lagrangian approach,
where the solution is based on moving frame of reference attached to the elements of
the moving mass (Hungr 1995). It is described as universal in that a broad range of
rheological models can be used within the program. The DAN program can also take
into consideration three-dimensional effects, such as for a confined flow, from
estimation of the slide width from the slope topography, and various rheological models
can be incorporated into the one travel path. Hungr (1995) gives a more detailed
discussion of the DAN program and its capabilities, and Fell et al (2000) and Hungr
(1995) give discussions on the various rheological models that are incorporated within
DAN.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.153

Table 3.29: Mean and standard deviation of H/L for several types of slopes giving “rapid” landslides.
Initial Slide Material Type / Degree of No. Volume H/L H/L H/L Std. Range of *3 Comments
Classification Slope Type Confinement of Cases Range Range Mean Dev. Std. Dev. of
Travel Path (cu.m.) Population
Confined, high 7 110,000 to 0.18 to 0.28 0.208 0.035 0.023 to 0.077 High mobility associated
mobility 5.6 x 10 6 with liquefaction
Coal mine waste susceptible materials
spoil piles (sandy mantling the down-slope.
2 No confinement 47 3,000 to 0.22 to 0.49 0.359 0.076 0.063 to 0.095 Similar mean and std. dev.
Flow Slides in gravels) *
condition, normal 8 x 106 for all travel path types.
Contractive
Soils mobility
Loose silty sand fills All unconfined 16 50 to 10,400 0.28 to 0.60 0.405 0.094 0.070 to 0.146
(Hong Kong)
Coking coal All unconfined 9 850 to 16,000 0.14 to 0.34 0.257 0.0625 0.042 to 0.120 Limited number of cases
stockpiles for analysis. Low specific
gravity
Natural Slopes (incl. Confined 19 50 to 13,000 0.22 to 0.67 0.426 0.110 0.083 to 0.162
For preliminary estimate of
Corominas debris Partly confined 10 80 to 3,000 0.28 to 0.62 0.470 0.114 0.078 to 0.207
travel distance angle
flow data) Unconfined 52 20 to 140,000 0.22 to 0.75 0.547 0.137 0.115 to 0.170
Failures in
Washout failures of Confined (assumed) 19 40 to 4,000 0.18 to 0.67 0.450 0.109 0.083 to 0.162 Silty sands of low clay
Dilative content.
silty sand fills
Soils*1
(Hong Kong)
Fills of silty sands Unconfined 10 40 to 500 0.52 to 0.84 0.641 0.093 0.064 to 0.170 Silty sands of low clay
(Hong Kong) content.
*1 Inclusive of defect controlled slides, slides of debris and slides through the soil mass.
*2 High mobility events associated with confined travel path and liquefaction susceptible materials mantling down-slope travel path.
*3 Statistical estimate of 95% confidence intervals of population based on case studies representing a sample of the population.
Std. Dev. = standard deviation.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.154

Table 3.30: Summary of recommended methods for prediction of H/L (tangent of the travel distance angle)
Initial Slide Slope / Material Properties Volume Degree of Method of Prediction of Reference Comments
Classification Material Range Confinement Travel Distance Angle Table / Figure
Type (cu.m.) / Equation
Coal mine Confined –high For curved down-slope Table 3.29, Consider confined high mobility
mobility angle use mean and Table 3.27 separately*1.
waste spoil Sandy gravels with low Up to
standard deviation. For normal mobility disregard
piles on fines content 5 x 106 Confined to Table 3.29,
If relatively constant use degree of confinement when using
Flow Slides in unconfined Table 3.27,
hillsides. Down-slope angle, a2 tables and figure.
Contractile Figure 3.75
Silty sands with 10 to 50 to 500 Unconfined Mean and standard Table 3.29
Soils
50% fines and low clay deviation
Case studies from flow slide failures
Fill Slopes 500 to Unconfined Mean and standard Table 3.29,
content (< 10%). Dry in loose fill slopes, Hong Kong
10,000 deviation, or Figure 3.73,
density < 85% SMDD.
slide volume Equation 3.18
Silty sands to sandy silts, Up to 100 Unconfined Cut slope angle, acut Equation 3.13 Derived from case studies of cut
low clay content, varying to 500 slope failures with run-out onto near
horizontal slopes at the toe.
gravel content (residual Up to Unconfined Cut slope angle, acut, and Figure 3.65, Lower bound in Figure 3.65 for
Cut Slopes
soils, saprolite and 20,000 slide volume Equation 3.14 Type 2 cut slopes.
colluvium) Use natural slope correlations if the
Failures in difference between the cut slope and
Dilative slope below is < 10 to 20 degrees.
Soils*1 Colluvial and some Confined Down-slope angle, a2 Figure 3.71, Use mean and standard deviation for
residual soils. Silty Equation 3.17 preliminary estimate of travel
Natural distance angle to establish a2 (from
sands to gravelly and Up to Partly confined Down-slope angle, a2 Figure 3.70, Table 3.29)
Slopes cobbly soils. Low clay 10,000 Equation 3.16 Then use Figure 3.69 to Figure 3.71
content. Unconfined Down-slope angle, a2 Figure 3.69, (or Equations 3.15 to 3.17) for more
Equation 3.15 accurate prediction.
*1 High mobility flow slides associated with suspected liquefaction of materials mantling the confined travel path. Normal mobility is not associated with the
liquefaction of materials mantling the travel path.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.155

3.10.1 “RAPID ” LANDSLIDING IN HONG KONG

Ayotte and Hungr (1998) and Hungr Geotechnical Research (1998) undertook
numerical modelling of some twenty-six reported “rapid” slides from Hong Kong using
the DAN program. The case studies analysed (Table 3.31) included 21 dilatant slides in
natural (and some quasi-natural) slopes, 4 failures in cut slopes (3 defect-controlled) and
1 flow slide failure in a loose fill slope. The debris flow travel classification would be
appropriate to all failures in natural terrain and cut slopes.
The purpose of the modelling was to attempt to match the actual debris deposition,
travel distance and velocity of the slide mass (where estimation was possible) using a
consistent rheological framework that could then be applied to similar potential
landslides within a risk assessment framework. The models used were the friction
model for unconfined and partly confined travel paths and dominantly translational
sliding type debris flows (e.g. Fei Tsui Road), and the Voellmy model for confined
debris flows. A combination of models was used for several landslides where the initial
travel path was on open or partly confined slope (using the frictional model) and the
distal part confined within a gully (using the Voellmy model).
The frictional model incorporates a frictional resistance component at the base of the
sliding mass (bulk friction angle, φb ) estimated from the normal effective friction ( φ ′ )
of the loose debris and an estimated pore pressure coefficient, ru, according to Equation
3.19.

tan φ b = ( 1 − ru ) tan φ ′ (3.19)

where r u = u / σ n , the ratio of the excess pore pressure, u, to the total normal stress, σ n .

The Voellmy model is a combination of frictional sliding and turbulent flow.


The findings of the analysis are summarised in Table 3.32. Based on the results,
Ayotte and Hungr (1998) concluded that:
• Good results were achieved with respect to matching the debris deposition profile,
travel distance and velocity (where available) of the back-analysis to the actual
event.
• For confined debris flows the Voellmy model, using a bulk friction angle of 5.7 to
11.3 degrees and turbulent friction coefficient of 500 m/sec2 resulted in a good
prediction of the slide mass travel. For predictive purposes, Ayotte and Hungr
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.156

(1998) suggest using a bulk friction angle of 11.3 degrees and turbulent friction
coefficient of 500 m/sec2 for the confined portion of debris flows (in Hong Kong).
• For unconfined and partly confined debris flows in natural slopes a broad variation
in the bulk friction angle resulted from the back-analysis (Figure 3.80), particularly
for small volume slides. Ayotte and Hungr (1998) considered the reasons for this to
be small obstructions in the flow path such as trees and boulders that affect the flow
behaviour of small magnitude slides, but not the larger magnitude slides. A
reasonable correlation is evident between the bulk friction angle of the travel path
and slide volume (Figure 3.80). They recommend the use of this correlation for
estimation of the bulk friction angle in the prediction of travel behaviour for open
slope failures of similar type and magnitude.
• For the larger volume events (greater than about 1000 m3) on open or partly
confined slopes the bulk friction angle is in the range 19 to 23 degrees. The findings
indicate that the travel of these larger volume mostly debris flows is independent of
the initiating mechanism (i.e. whether it is a flow slides, defect-controlled failure or
slide of debris) and slope type (cut, fill or natural slope). Although, it is difficult to
draw any definitive conclusions from the small number of slides analysed larger
than about 1000 m3.

Table 3.31: Summary of “rapid” landslides from Hong Kong analysed by Hungr
Geotechnical Research (1998) and Ayotte and Hungr (1998)
Slide No. Volume Initial Slide Source Area Travel
Type Cases Range (m3) Classification Material Classification
Description
Cut 4 85 to 40000 3 defect-controlled 3 in volcanic 4 debris flows. Rear
(3 greater 1 dilatant slide saprolite, 1 in portion of Po Shan
through saprolitic granitic saprolite (1972) possibly a debris
than 2000 m3)
soil mass. slide.
Fill 1 2500 Flow slide in loose, Silty sand Flow slide
end dumped fill (decomposed granite)
Natural 21 (2 50 to 26000 17 slides of debris Slides of debris in 4 confined debris flows.
quasi (most less 1 defect-controlled colluvium mantling 9 partly confined debris
3 dilatant slides the slope. flows.
natural) than 1000)
through saprolite or Of the 4 remaining 8 debris flows
residual soil mass. slides, 1 was in (unconfined), 1 of
residual volcanic, 2 which is a debris slide /
were in volcanic debris flow.
saprolite, and 1 was
in granitic saprolite.
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.157

Ayotte and Hungr (1998) encountered difficulties modelling several slides within the
framework of using a consistent rheologic approach. Three of the partly confined debris
flows were modelled using the Voellmy model. In addition, they had difficulties getting
the debris to initially slide out of the source area and overcame this problem by using a
lower bulk friction angle for the source area.
The back-analysis results (Table 3.32 and Figure 3.80) present useful guidelines for
selection of model type and material parameters for use in modelling the travel of slides
in similar soil types based on slide type, volume and degree of confinement of the travel
path. It is interesting that for the larger volume unconfined and partly confined events
the mechanism of failure (flow slide, slide of debris or defect-control) has a limited
influence on the estimated bulk friction angle. The limited number of case studies of
each failure mechanism type makes it difficult to draw any firm conclusions as to the
reason for this finding; however, it is considered possible that it reflects the large degree
of break-up and flow type behaviour of the slide mass. Fei Tsui Road stands out as an
outlier and this possibly reflects the lesser degree of break-up and suspected significant
basal sliding component of the slide mass. For the small volume slides small
obstructions to the flow are a dominant factor in the debris travel behaviour (as
concluded from the empirical analysis) that is difficult to take into consideration and
significant uncertainty is prevalent in predictive modelling.

Table 3.32: Summary of results of numerical modelling using DAN for “rapid”
landslides in Hong Kong (Hungr Geotechnical Research 1998; Ayotte and Hungr 1998)
Slide Type Source No. Frictional Model Voellmy Model
Volume Cases φb source φb path φ (º) Turbulent
Range (m3) (º) (º) Coefft
Natural Slopes – <200 4 14 to 30 24 to 43 - -
unconfined and partly 200 to 500 6 23 to 34 25 to 34
confined using frictional 500 to 26000 5 20 to 25 20 to 25
model.
Natural Slopes – partly 150 to 450 3 11 to 500
confined using Voellmy 22
model
Natural Slopes - 50 to 13000 4 - - 5.7 to 500
confined 11
Cut Slopes – defect 85 to 40000 4 14 to 24 20, 23, 26 - -
controlled & 36
Flow Slide in loose fill 2500 1 20 20 - -
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.158

1.0

Tan (bulk friction angle) of travel path . 0.9

0.8
beyond toe of source area .

Fei Tsui Road


0.7

0.6

0.5

0.4

0.3
Natural slopes - unconfined debris flows
0.2
Natural slopes - partly confined debris flows
Cut slopes
0.1
Fill slopes - flow slide in loose fill
0.0
1.E+01 1.E+02 1.E+03 1.E+04 1.E+05
Volume (cu.m.)

Figure 3.80: DAN results of frictional back-analysis models of “rapid” landslides from
Hong Kong (data from Hungr GR (1998) and Ayotte and Hungr (1998))

3.10.2 “RAPID ” LANDSLIDES FROM COAL MINE WASTE SPOIL PILE FAILURES IN
BRITISH COLUMBIA

Golder Assoc. (1995) undertook numerical modelling of some forty-one case studies of
“rapid” flow slides from failures in coal mine waste spoil piles from British Columbia
using the DAN program. The purpose of the modelling (as for Hong Kong) was to
attempt to match the actual debris deposition, travel distance and slide mass velocity
(where estimation was possible) using a consistent rheological framework that could
then be applied to similar potential landslides within a risk assessment framework. The
models used were the frictional model for unconfined and partly confined travel paths
(generally less than 1 km travel distance) and the Voellmy model for confined flow
slides. A combination of models was used for several “rapid” landslides where the
initial travel path was unconfined (using frictional model) and the distal portion
confined within a gully (using Voellmy model).
Based on the findings of the numerical analyses, Golder Assoc. (1995) concluded
that:
• A good simulation of the observed velocities, debris deposition profile and travel
distance was achieved for the partly confined and unconfined flow slides with travel
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.159

distances less than 1 km using the frictional model with bulk friction angles of 18 to
24 degrees (Figure 3.81). For an estimated dynamic friction angle of 30 degrees for
the debris, this equates to pore pressure coefficients in the range 0.23 to 0.44.
• The frictional model was less suitable for the confined flows with travel distances in
excess of 1 km. Quite low bulk friction angles (10 to 18 degrees) were required to
match the travel distance and the modelled velocities were up to three times greater
than field estimates.
• Using the Voellmy model for the portion of the confined travel path for the large
travel distance cases provided a reasonable simulation of the debris deposition and
velocities when matching the travel distance. The parameters used for the Voellmy
model were a bulk friction angle of 3 to 6 degrees and turbulent friction coefficient
of 200 to 300 m/sec2.

Based on the results from Golder Assoc. (1995), it is concluded that for predictive
modelling purposes it is necessary to calibrate against known failures in similar
materials, with similar degree of confinement and foundation conditions along the travel
path. Otherwise it is necessary to accept large uncertainty in the use of numerical
models.

Figure 3.81: Back calculated bulk friction angle from DAN analysis of flow slides from
coal mine waste spoil pile failures in British Columbia (Golder Assoc. 1995)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.160

3.10.3 FLOW SLIDES FROM TAILINGS DAMS FAILURES

Numerical analysis has been undertaken for flow slides from tailings dam failures by
Jeyapalan et al (1983b) and Blight (1997).
Blight (1997) undertook a simple sliding block model analysis (Figure 3.82)
assuming hydrostatic pore pressures within the flowing tailings and shear resistance to
flow along the interface between the down-slope foundation and flowing tailings. He
used the model to estimate the shear resistance along the interface (Equation 3.20) from
the known characteristics of the slide.

τ =
(P1 − P2 )cos i + W sin i (3.20)
L / cos i

From analysis of 8 case studies Blight (1997) concluded that the undrained residual
strength of the liquefied tailings ranged from 1 to 4 kPa and that the shear strength was
sensitive to the availability of free water down-slope of the failure. For the cases that
flowed onto a wet down-slope (from recent rainfall) the shear resistance along the
interface ranged from 1.0 to 2.6 kPa. Where the tailings flowed onto a dry down-slope
the shear resistance along the interface ranged from 3.4 to 5.2 kPa. From laboratory
testing (Figure 3.83) it is evident that the shear strength of liquefied tailings (derived
from the crushing of ore) decreases rapidly with increasing moisture content.
Therefore, any free water entrained within the base of a laminar type flow of tailings
can significantly reduce the shear strength.

Figure 3.82: Simple sliding block model for analysis of tailings strength (Blight 1997)
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.161

Figure 3.83: Variation of shear strength with moisture content for gold tailings (Blight
1997)

Jeyapalan et al (1983a) and Blight (1997) comment that the Bingham plastic rheologic
model is a good approximation to the flow of liquefied tailings where the flow is
laminar. For turbulent flow of tailings, i.e. a hyper-concentrated stream-flow, Jeyapalan
et al (1983a) recommend the use of flood routing computer programs to model the flow.
Jeyapalan et al (1983a) developed solutions using the Bingham plastic rheological
model to a dam break type analysis (Figure 3.84) for flow on horizontal sloping planes
(one-dimensional model) and for flows within prismatic valleys. They developed
graphical solutions for the horizontal sloping plane analysis based on the strength and
viscous parameters of the flow for down-slope angles ranging from 0 to 14 degrees.
Jeyapalan et al (1983b) tested the model against flume tests using viscous oil and
found good agreement between the model and the test. They then applied the model to
two actual failures (Aberfan in 1966, which is not a tailings dam failure, and the
gypsum tailings dam failure in Texas in 1966) and reported reasonable agreement even
though the model is limited to one-dimension and therefore cannot take into
consideration spreading of the flow. Golder Assoc. (1995) found that the Bingham
plastic rheologic model was not suitable for flow slides in the coal waste spoil piles of
British Columbia.
The degree of agreement between the case study and the numerical model for the
Bingham plastic rheologic model is very strongly dependent on the estimated value of
viscosity and yield strength. Given the potential broad variation in shear strength of
tailings (Figure 3.83) use of the model for predictive purposes should be undertaken
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.162

cautiously and factors such as potential entrainment of water in the slide mass should be
considered. Its use is likely to be more applicable to laminar type flow conditions of
fine grained tailings flows as the findings by Jeyapalan et al (1983b) and Golder Assoc.
(1995) suggest.
Jeyapalan et al (1983b) also applied a turbulent flow model (using a flood routing
program) to the catastrophic failure at Buffalo Creek. They reported comparable results
between the model and the actual flow. This type of flow behaviour (hyper-
concentrated stream flow) is not uncommon within tailings dam failures. Other cases of
turbulent or hyper-concentrated stream flows are considered to include the confined
flow portions of the tailings dam failures at Stava, Italy (in 1985), Bafokeng, South
Africa (in 1974) and a number of the earthquake induced failures that flowed for many
kilometres.

Figure 3.84: Model for flow of liquefied tailings (Jeyapalan et al 1983a)

3.11 CONCLUSIONS
“Rapid” landslides from failures in soil slopes have the potential to cause mass
destruction, resulting in loss of life, destruction to property and damage to the natural
environment. It is important therefore that geotechnical engineers, engineering
geologists and geologists understand the soil characteristics, slope conditions and failure
mechanics that can result in “rapid” sliding, and are alert to the pre-failure warning
signs prevalent for several classes of slope failure that develop into “rapid” landslides
post failure. An important part of any risk assessment of potential slope instability is
prediction of the post-failure deformation behaviour for evaluation of the risk to life,
property and the environment down-slope (or up-slope in the case of retrogressive
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.163

failures) of the potential failure. An understanding of the soil characteristics (both the
slide mass and slope mantling materials), failure mechanics, down-slope conditions and
travel behaviour is important in making informed predictions of travel distance.
The mechanics of failure of landslides that develop into “rapid” slides has been
divided into two broad classifications; flow slides in contractive soils and landslides in
dilative soils. Several classes of landslide have been identified and analysed within
each classification, they are:
• For flow slides in contractile soils:
− Flow slides (sub-aerial) in loose dumped fills of soil types ranging from silty
sands to hillside mine waste spoil piles of sandy gravels.
− Flow slides in hydraulically placed fills (including tailings dams, hydraulic fill
embankment dams and sub-aqueous fills).
− Flow slides in natural soils (including sensitive clays and submarine slopes)
• For landslides in dilative soils, failures within steep cut, fill and natural slopes have
been analysed. Two dominant failure types have been identified:
− Strongly defect-controlled slope failures in weathered soil slopes (residual soil
and weathered rock masses) along discontinuities adversely orientated to the
face slope.
− Slope failures within the soil mass (with no defect control) including; slides of
debris in colluvium and failures through the soil mass in residual soils,
weathered rock masses and compacted fills.

An important component of the research was the literature review on the mechanics
of shearing of contractive, granular soils. The characteristics of soils susceptible to
static liquefaction and potential flow sliding, and methods of assessment of flow-
liquefaction potential are summarised in Section 3.6. Section 3.3.5 summarises the
available methods for assessment of the residual undrained strength of flow-liquefied
soils. Section 3.4 provides additional information on the soil properties, triggering and
failure mechanics of flow slides for the categories of flow slide considered.
For failures in dilative soils, Section 3.6 summarises the material properties and
slope characteristics within which the failure has developed into a “rapid” debris flow or
debris slide post failure. Additional information is provided in Section 3.5, for both
Chapter 3: “Rapid” landslides from failures in soil slopes Page 3.164

failures through the soil mass (Section 3.5.2) and strongly defect-controlled failures
(Section 3.5.3).
Travel distance angle, and therefore travel distance, of the failed slide mass can be
predicted empirically using correlations with slide volume, down-slope angle and other
factors. Section 3.9.5 summarises the analysis on methods of travel distance angle
prediction and Table 3.30 (within Section 3.9.5) summarises the recommended methods
for prediction of the travel distance angle for various slope and material types. These
methods are, in most cases, improvements on current methods. Details of the analysis,
findings and limitations of the predictive methods for the various slope and material
types analysed in detail are given in Sections 3.9.3 and 3.9.4.
There is inherently a large uncertainty associated with predictions of post-failure
deformation analysis, for both empirical and numerical analyses. For important projects
it is recommended that empirical methods and numerical models be calibrated using
local, similar type failures to reduce some of the uncertainty. Risk analysis methods can
also be incorporated into the analysis by assessment of the mean and range of outcomes.
Chapter 4: Embankments on soft ground Page i

TABLE OF CONTENTS

4.0 EMBANKMENTS ON SOFT GROUND................................... 4.1


4.1 INTRODUCTION TO THIS C HAPTER ................................................................... 4.1
4.2 LITERATURE R EVIEW ...................................................................................... 4.3
4.2.1 Excess Pore Water Pressure Response ..................................................... 4.5
4.2.2 Deformation Behaviour...........................................................................4.15
4.3 ANALYSIS OF THE B EHAVIOUR OF FAILURE CASE STUDIES ........................... 4.22
4.3.1 Case Studies Analysed............................................................................. 4.22
4.3.2 Excess Pore Water Pressure ................................................................... 4.26
4.3.3 Pre-Failure Deformation Behaviour and the Effects of Progressive
Failure..................................................................................................... 4.32
4.3.4 Factor of Safety Versus Relative Embankment Height............................ 4.45
4.3.5 Post-Failure Deformation Behaviour ..................................................... 4.47
4.4 POST CONSTRUCTION BEHAVIOUR OF EMBANKMENTS ON SOFT GROUND .... 4.51
4.4.1 Effect of Effective Stress State on the Post Construction Behaviour....... 4.52
4.4.2 Deformation Behaviour and Pore Water Pressure Response in the
Initial Period Post Construction ............................................................. 4.52
4.4.3 Long-Term Post Construction Deformation............................................ 4.59
4.5 DISCUSSION AND CONCLUSIONS.................................................................... 4.61
4.5.1 Indicators of an Impending Failure ........................................................ 4.61
4.5.2 Guidelines on Monitoring for Identification of an Impending Failure
Condition................................................................................................. 4.63
4.5.3 Post Failure Deformation Behaviour...................................................... 4.64
4.5.4 Discussion on Failure Mechanism.......................................................... 4.65
Embankments on soft ground Page ii

LIST OF TABLES
Table 4.1: Summary of failure case studies of embankments on soft ground. ............ 4.24
Table 4.2: Excess pore water pressure observations for the failure case studies......... 4.26
Table 4.3: Summary of exceptions to general observations of excess pore water
pressure response. ................................................................................................. 4.31
Table 4.4: Summary of case studies that are exceptions or have other explanations
to deformation as an indicator of impending failure............................................. 4.44
Table 4.5: Results of the post-failure deformation analysis......................................... 4.50

LIST OF FIGURES
Figure 4.1: Computed factor of safety for field failures (Tavenas and Leroueil
1980) ....................................................................................................................... 4.2
Figure 4.2: Effective vertical stress profile and pre-consolidation stress profile
prior to construction, and the effective vertical stress profile at end of
construction of the Saint Alban B test embankment (Tavenas and Leroueil
1980). ...................................................................................................................... 4.8
Figure 4.3: Field observations of excess pore water pressures in clay foundations
during the initial period of embankment construction (Tavenas and Leroueil
1980). ...................................................................................................................... 4.8
Figure 4.4: Total (OPFR) and effective (O′P′F′R′) stress paths under the centre of
embankments (Tavenas and Leroueil 1980) ........................................................... 4.9
Figure 4.5: Relationship between the excess pore water pressure and the applied
vertical stress (Tavenas and Leroueil 1980)............................................................ 4.9
Figure 4.6: Dimensionless variations of changes in total stresses under the
centreline of an infinite strip load on an elastic half space (Poulos and Davis
1974) ..................................................................................................................... 4.10
Figure 4.7: Effective stress paths for varying B values and total stress ratios
(Folkes and Crooks 1985). .................................................................................... 4.11
Figure 4.8: Effective stress paths and pore water pressure generation associated
with different types of yield conditions (Leroueil and Tavenas 1986). ................ 4.11
Figure 4.9: Element of soil at depth z beneath on centreline of uniform circular
load (Muir Wood 1990) ........................................................................................ 4.13
Figure 4.10: Total stresses on centreline beneath uniform circular load on elastic
half space (Muir Wood 1990). .............................................................................. 4.14
Embankments on soft ground Page iii

Figure 4.11: Total (BCDE) and effective (B′C′D′E′) stress paths for a soil element
on the centreline beneath a circular load (Muir Wood 1990). .............................. 4.14
Figure 4.12: Variation of pore water pressure with applied surface load for
element on centreline beneath circular load (Muir Wood 1990). ......................... 4.15
Figure 4.13: Variation of the dimensional factor R (proportional to the horizontal
surface displacement at the toe) with factor of safety (Marche and Chapuis
1974) ..................................................................................................................... 4.17
Figure 4.14: Definition of the geometry and deformation parameters (Tavenas et
al 1979).................................................................................................................. 4.18
Figure 4.15: Average correlation between the maximum lateral movement at the
toe, ym, and settlement under the centre of the embankment, s, during
construction (Tavenas and Leroueil 1980)............................................................ 4.18
Figure 4.16: Effective stress and lateral displacement profiles at the end of
construction under (a) Cubzac-les-Ponts A and (b) Saint Alban B test fills
(Tavenas et al 1979) .............................................................................................. 4.19
Figure 4.17: Comparison of horizontal movement (on berm beyond embankment
toe) during loading and for 24 hour period for the Thames test embankment
(Marsland and Powell 1977) ................................................................................. 4.20
Figure 4.18: Observations of lateral displacement at the base of the embankment
with increasing embankment thickness at the Rio test embankment (Ramalho-
Ortigao et al 1983a)............................................................................................... 4.21
Figure 4.19: Contours of normalised shear-stress ratios at 4 m fill thickness from
finite element modelling at Muar-F test embankment, Malaysia (Indraratna et
al 1992).................................................................................................................. 4.22
Figure 4.20: Idealised section of embankment indicating the types of monitored
observations........................................................................................................... 4.23
Figure 4.21: (a) and (b) Excess pore water pressure under centre of embankment
versus applied embankment load. ......................................................................... 4.27
Figure 4.22: Settlement under the centre of the embankment versus relative
embankment height (H/Hf). ...................................................................................4.34
Figure 4.23: Vertical displacement at the embankment toe versus relative
embankment height. .............................................................................................. 4.35
Figure 4.24: Vertical displacement beyond the embankment toe versus relative
embankment height. .............................................................................................. 4.36
Figure 4.25: Lateral surface displacement at the embankment toe versus relative
embankment height. .............................................................................................. 4.38
Embankments on soft ground Page iv

Figure 4.26: Incremental vertical inclination angle (from inclinometer


observations at the embankment toe) with fill height (and time) for (a)
Thames and (b) King’s Lynn. ............................................................................... 4.39
Figure 4.27: Vertical inclination angle on the surface of rupture from inclinometer
observations at the embankment toe versus relative embankment height. ........... 4.40
Figure 4.28: Maximum lateral displacement at the embankment toe versus
settlement under the centre of the embankment. ................................................... 4.43
Figure 4.29: Factor of safety versus relative embankment height ............................... 4.46
Figure 4.30: Failure model for post-failure deformation (Khalili et al 1996).............. 4.49
Figure 4.31: Effective stress paths post construction (Leroueil and Tavenas 1986).... 4.53
Figure 4.32: Excess pore water pressure response with time for several
piezometers at the Muar-EC test embankment. .................................................... 4.55
Figure 4.33: Excess pore water pressure response at Piezometer C1 with time at
the St. Alban-A test embankment. ........................................................................ 4.56
Figure 4.34: Schematic behaviour of strain rate effects on limit state surface with
decreasing strain rate in undrained creep tests (adapted from Leroueil and
Marques 1996) ...................................................................................................... 4.58
Figure 4.35: Relationship between the maximum lateral displacement at the toe,
ym, and settlement at the centre, s, at the end of construction (Tavenas et al
1979). .................................................................................................................... 4.60
Figure 4.36: Relationship between maximum lateral displacement (toe) and
settlement (centre) at the end of construction for different embankment
geometries and stability conditions (Tavenas et al 1979). .................................... 4.61
Chapter 4: Embankments on Soft Ground Page 4.1

4.0 EMBANKMENTS ON SOFT GROUND

4.1 INTRODUCTION TO THIS CHAPTER


The results of stability analysis of failure case studies of embankments on soft ground
(Bjerrum 1972; Pilot 1972; Aas et al 1986) indicate conventional total stress stability
analysis is unreliable in prediction of the failure height, with failure generally occurring
at factors of safety of greater than unity. The uncertainty in the assessment of the
stability of embankments on soft ground is due to the difficulty in predicting the
undrained shear strength of the soils in the foundation, given the inherent variability of
the soil strata and the uncertainty in the methods of measuring the strength; modelling
the strength of the upper over-consolidated, often fissured, desiccated crust; the
potential for fissuring in the lightly over-consolidated soils below the crust; and the
potential for cracking in the embankment.
To account for this uncertainty various guidelines have been developed for stability
analysis of embankments on soft ground and methods of interpretation of field data.
Bjerrum (1972) proposed a correction factor based on the plasticity of the soft clay
foundation to account for the discrepancy between the vane strength and the mobilised
strength at failure. The predominant factors attributing to the strength discrepancy were
considered by Bjerrum (1972, 1973) to be the difference in the rate of loading between
the vane and actual embankment construction, strength anisotropy and progressive
failure. As shown in Figure 4.1 there is significant uncertainty surrounding this
correction factor. Other guidelines included; correction factors for specific soil types,
such as leached, sensitive marine deposits (Dascal and Tournier 1975); methods of
assigning strength properties to fissured soils, particularly the crust (La Rochelle et al
1974; Ferkh and Fell 1994); the effect of plant matter and shells on the vane strength
(La Rochelle et al 1974; Ferkh and Fell 1994; Marsland and Powell 1977); and
treatment of the strength of the embankment.
However, even with the methods of interpretation and analysis available, the Type A
predictions for the Muar test embankment (Brand and Premchitt 1989) highlight the
uncertainty that is still prevalent in limit equilibrium analysis of embankments on soft
ground. Of thirty-one predictions, the predictions of fill height at failure ranged from
2.8 m to 9.5 m (mean = 4.7 m, standard deviation = 1.5 m, actual = 5.4 m) and the
predictions of the depth of the failure surface ranged from 2.5 m to 13 m depth (mean =
Chapter 4: Embankments on Soft Ground Page 4.2

6.3 m, standard deviation = 2.2 m, actual = 8 m). Most of this variation may relate to
those involved not being directly responsible for the project, and failure to differentiate
between designing a safe embankment and predicting a failure height, but there is a
significant element of uncertainty inherent at this time.
El-Ramley (2001) showed that for the soft ground case studies he studied, a factor
of safety of 1.5 corresponded to a probability of unsatisfactory performance (failure) of
about 1 in 50. This reflected the large uncertainty in estimating the undrained strength.
The uncertainty surrounding the estimation of factor of safety (and settlement
behaviour) for embankments on soft ground has led to the construction of full-scale test
embankments, particularly for large projects. In a number of cases these test
embankments (which were generally well-monitored) were constructed to failure to
determine appropriate design and construction procedures of the actual embankment.

Figure 4.1: Computed factor of safety for field failures (Tavenas and Leroueil 1980)

The main purpose of the research on embankments on soft ground was to evaluate the
pre-failure deformation behaviour and pore water pressure response from well-
monitored case studies constructed to failure, and to identify field indicators of an
impending failure condition. This will provide designers and construction supervisors
with guidelines on interpretation of monitoring allowing early identification of an
impending failure condition, which would allow appropriate measures to be undertaken
Chapter 4: Embankments on Soft Ground Page 4.3

to avert failure (eg. stopping construction to allow pore water pressures induced by the
embankment construction to dissipate, and the foundation to gain strength).
A secondary objective of the research was to develop further the understanding of
post failure deformation and evaluate the applicability of existing models to predict the
post failure deformation behaviour.
Initially, a summary is given of the literature in the context of the factors and
findings affecting the deformation behaviour of embankments on soft ground (Section
4.2). It is by no means exhaustive, but focuses on observations, models and proposed
methods related to the mechanics controlling the deformation, and causing changes in
the deformation behaviour during loading, and the indicators of deformation behaviour
leading up to a failure condition. Thirteen case studies of well-documented and
monitored embankments on soft ground that were constructed to failure are then
analysed (Section 4.3). Guidelines are developed for identification of an impending
failure condition taking into consideration the mechanics of failure and the soil
properties (Section 4.5). The general trend of the deformation behaviour and pore water
pressure response is described within this framework.
The analysis of the post failure deformation behaviour of four of the embankment
on soft ground case studies was undertaken using the Khalili et al (1996) method for
intact sliding (Section 4.3.5.2). General guidelines on the applicability and use of this
type of model are given in Section 4.5 and are discussed in Chapter 5 together with the
findings from analysis of failures in embankments dams and cut slopes in high plasticity
heavily over-consolidated clays (Sections 5.8.1 and 5.9.2).

4.2 LITERATURE R EVIEW


This literature review concentrates on developing an understanding of the mechanics of
the deformation behaviour during construction and leading up to failure. Predictive
methods from the literature and further discussion of the factors affecting the
deformation behaviour are presented with the findings of the case study analyses in
Section 4.3. Related issues such as stability and consolidation behaviour are not
discussed in any detail, but are implicit in the analysis of the data.
Bjerrum (1972, 1973), in his review of settlement behaviour of embankments on
soft ground, recognised (as had Höeg et al 1969; and D’Appolonia et al 1971a) that the
Chapter 4: Embankments on Soft Ground Page 4.4

in-situ over-consolidated state of the soft clay foundation had a significant influence on
the pore water pressure response and settlement behaviour. Bjerrum commented that:
• At stress levels less than the pre-consolidation pressure, settlements are small and
dissipation of pore water pressure is rapid. He described this as being due to the
over-consolidated clays having high coefficients of consolidation, cv, as a
consequence of their low compressibility (related to the recompression index, Cs)
and relatively small amount of pore water required to flow out of the soil structure.
• At stress levels in excess of the pre-consolidation pressure (normally consolidated
condition) total settlements (during and post construction) are relatively large due to
the high compressibility of the clay. The initial excess pore water pressures are high
and dissipate relatively slowly due a much lower coefficient of consolidation, cv, of
the normally consolidated clay. The “immediate” settlement during loading occurs
as a result of lateral straining associated with plastic yielding of the normally
consolidated foundation.

The one-dimensional relationship between the soil stiffness (Cc, compression index) and
coefficient of consolidation, cv, proposed by Terzaghi is:


k kσ v (1 + e )
cv = = (4.1)
γ w mv 0.434γ wC c

where k is the hydraulic conductivity, e the void ratio, γ w the unit weight of water,

σ v ′ the effective vertical stress and m v the coefficient of volume change. Lambe and

Whitman (1979) report ratios of the compression index, Cc, for clays in a normally
consolidated condition compared to similar clays in an over-consolidated condition in
the range 3 to 50, indicating the significant effect on cv of compressibility between
stress levels below and above the pre-consolidation pressure.
Höeg et al (1969) described the change between the two states as a critical value of
applied stress at which settlements increased rapidly and a pronounced increase in rate
of pore water pressure build up occurred with additional loading. D’Appolonia et al
(1971a) also observed a similar increase in rate of pore water pressure development per
unit increase in load in the MIT test embankment and attributed this increase to local
yielding and contained plastic flow.
Chapter 4: Embankments on Soft Ground Page 4.5

4.2.1 EXCESS PORE WATER PRESSURE RESPONSE

The Geotechnical Section at Université Laval, Quebec and the Research Group on
Compressible Soils at the Laboratoire Central des Ponts et Chaussées, Paris undertook a
joint research effort in the 1970’s to develop further the understanding of deformation
behaviour in soft clays under embankments. The following is a summary of the
pertinent findings of their research related to field observation of excess pore water
pressure response within the foundation during embankment construction. Discussion
on the findings of their research on the deformation behaviour of embankments on soft
ground is given in Section 4.2.2.
Leroueil et al (1978a) analysed the pore water pressure response observed under the
centre of 30 embankments on soft clay from case studies covering a wide variety of
clays in the Americas, Europe and Asia. The observed behaviour was interpreted using
the YLIGHT model of clay behaviour (Cam Clay model incorporating strength
anisotropy) proposed by Tavenas and Leroueil (1977). They recognised that all clays
had developed some pre-consolidation (Figure 4.2), by aging, past loading, and other
mechanisms such as desiccation by surface drying or evapo-transpiration of plants, and
fluctuations in the ground water levels. For lightly over-consolidated clay (OCR < 2.5,
OCR = over-consolidation ratio) the observed behaviour and stress path interpretation is
summarised as follows:
i) In the early stages of loading (at effective stress levels less than the pre-
consolidation pressure, σ ′p ) the excess pore water pressures generated ( ∆u ) are

influenced by the process of consolidation, i.e. dissipation of pore water pressure


occurs during loading due to the relatively high coefficient of consolidation, cv.
Maximum excess pore water pressures developed in the central portion of the clay
layer corresponded to B (Equation 4.2) values of 0.38 to 0.75, and the pore water
pressure profile with depth was similar to that of a consolidation isochrone (Figure
4.3). The observed excess pore water pressure response was significantly below
that corresponding to an undrained condition. In this stage of loading the over-
consolidated clay foundation is characterised by relatively high stiffness (i.e. low
compressibility) and high cv, and dissipation of excess pore water pressures
therefore occurs during construction. Path O′P′ (Figure 4.4) is idealised as the
effective stress path for this stage of loading.
Chapter 4: Embankments on Soft Ground Page 4.6

∆u   ∆σ 3  
B= = B 1 − (1 − A)  1 −   (4.2)
∆σ 1   ∆σ 1 

ii) At some stage in the construction process, the excess pore water pressure generated
(at a specific piezometer) is observed to increase to an incremental B value of
approximately 1 (Figure 4.5). Leroueil et al (1978a) interpreted this to be at Point
P′ (Figure 4.4), where the effective vertical stress at the piezometer is equivalent to
the pre-consolidation pressure, i.e. the limit state surface defining the normally
consolidated state. The location is termed the “critical embankment load”, σ 1′ ,crit ,
or the “threshold embankment height”, Hnc.
iii) With further loading, the excess pore water pressure response is typically at an
incremental B value of approximately 1 ( B2 in Figure 4.5), i.e. an undrained
condition with the clay in a normally consolidated condition. In this stage of the
loading the clay is characterised by high compressibility and low cv resulting in low
rates of excess pore water pressure dissipation. In most of the cases analysed by
Leroueil et al (1978a), the rate of construction was sufficiently high that negligible
excess pore water pressure dissipation occurred. In terms of the interpreted stress
path (Figure 4.4) the effective stress path for this stage of loading is P'F'.
iv) With further loading, the effective stress path reaches the critical state surface
(Point F´ in Figure 4.4) and the soil reaches a localised failure condition. A
localised failure is generally initiated in the foundation below the shoulder of the
embankment where shear stresses are greatest. For:
− Strain weakening soils, further straining of a soil at Point F´ results in a
reduction in shear strength or deviator stress (along path F´R´) and therefore
generation of excess pore water pressures within the localised failure zone of
greater than the applied load ( B f > 1 , Figure 4.5). The total stress path shifts to

the right with a reducing deviatoric stress (path FR in Figure 4.4). Tavenas and
Leroueil (1980) consider that the initial structure of the soil is partly destroyed
during deformation in a normally consolidated condition, resulting in a change
in mechanical properties and modification of the shape of the limit state surface
from that originally represented by F´Yo to F´R´ (Figure 4.4).
− Soils that are not strain weakening, the effective stress state does not change
from Point F´. As discussed by Muir Wood (1990), application of additional
Chapter 4: Embankments on Soft Ground Page 4.7

load results in a sideways shift in the total stress path (Figure 4.11), with pore
water pressure increases that are equal to the change in total vertical stress
(Figure 4.12), and plastic deformation occurs.
v) Complete failure of the embankment does not necessarily occur once a localised
failure condition is reached. The localised failure zone is supported by the
surrounding unfailed soil.

Equations 4.3 and 4.4 (Tavenas and Leroueil 1980) define the excess pore water
pressure generated up to the threshold height (Equation 4.3) and beyond the threshold
height (Equation 4.4), as shown in Figure 4.5.

∆u = B1 IγH (4.3)

∆u = B 2 I∆ γH where B 2 = 1 (4.4)

where ∆u is the excess pore water pressure, I is the stress influence factor from elastic
solutions, γH is the applied load of the embankment and ∆γH is the change in the

applied load above the threshold height, Hnc. B1 and B2 are defined in Figure 4.5. The
threshold height is defined in Equation 4.5.

(σ ′ − σ ′ )
H nc =
p vo

γI (1 − B )
(4.5)
1

It is important to note that the threshold height varies with depth. From Equation 4.5,
the threshold height at any location is dependent on the difference between the pre-
consolidation pressure and initial effective stress (σ ′p − σ vo
′ ), the pore water pressure

response ( B1 ) and the stress influence factor; i.e. the shape of the effective vertical
stress profile and how this relates to the pre-consolidation stress profile. Figure 4.2
shows the zone of normally consolidated soft clay for an effective stress profile, σ ′v ,
determined from the applied total stress and pore water pressure response. With
increasing applied load the zone of normally consolidated clay will also increase. The
effect of this on the deformation behaviour is discussed in Section 4.2.2.
Chapter 4: Embankments on Soft Ground Page 4.8

Where:
z = depth below ground surface
σ ′vo = initial effective vertical stress
σ ′p = pre-consolidation stress
σ ′v = effective vertical stress at end of
construction

Figure 4.2: Effective vertical stress profile and pre-consolidation stress profile prior to
construction, and the effective vertical stress profile at end of construction of the Saint
Alban B test embankment (Tavenas and Leroueil 1980).

Where:
Z/D is the normalised depth with respect to
the depth of the soft clay (measured from
ground surface level).
B1 = B measured at effective stress levels
less than the pre-consolidation pressure.

Figure 4.3: Field observations of excess pore water pressures in clay foundations during
the initial period of embankment construction (Tavenas and Leroueil 1980).
Chapter 4: Embankments on Soft Ground Page 4.9

Figure 4.4: Total (OPFR) and effective (O′P′F′R′) stress paths under the centre of
embankments (Tavenas and Leroueil 1980)

Figure 4.5: Relationship between the excess pore water pressure and the applied vertical
stress (Tavenas and Leroueil 1980)

Whilst more recent publications concur with the pore water pressure response described
by Tavenas and Leroueil (1980), there is conjecture on the effective stress path,
particularly in the over-consolidated region. Folkes and Crooks (1985) comment that
B does not provide a satisfactory description of the effective stress path direction since
it is a function of the total stress ratio, ∆σ 3 / ∆σ 1 , (Equation 4.2) and the total stress

ratio varies with depth (Figure 4.6). As Figure 4.7 indicates, B does not provide an
indication of the direction of the effective stress path. Folkes and Crooks (1985)
Chapter 4: Embankments on Soft Ground Page 4.10

recommend that more emphasis be placed on the actual effective stress path rather than
the value of B .
Leroueil and Tavenas (1986) agree that B does not provide a satisfactory
description of the effective stress path, but still contend that partial drainage does occur
during the initial stages of embankment construction. Figure 4.8 presents a summary of
effective stress paths and pore water pressure response associated with different types of
yield conditions for embankments constructed to failure (Leroueil and Tavenas 1986).
In the context of the discussion to date on pore water pressure response Figure 4.8b is
the relevant effective stress path and is considered to be generally representative of the
effective stress path for most of the failure case studies analysed in Section 4.3. Figure
4.8a relates to case studies where limited drainage occurred during construction whilst
the effective stress state was below the pre-consolidation pressure (i.e. in an over-
consolidated state). Examples and discussion of this behaviour are given in Appendix C
(Section 3).

total vertical stress


0.5
total horizontal stress
deviator stress
1
Dimensionless Depth, z/b

total mean stress


total stress ratio
1.5

Note: (σ 1′ + σ ′3 ) / 2
2.5 total mean stress =

deviator stress =
(σ 1′ − σ 3′ )
3
z = depth below base of strip load
b = half width of the strip load
3.5

4
0 0.2 0.4 0.6 0.8 1
Dimensionless Total Stress

Figure 4.6: Dimensionless variations of changes in total stresses under the centreline of
an infinite strip load on an elastic half space (Poulos and Davis 1974)
Chapter 4: Embankments on Soft Ground Page 4.11

Note: p ′ = (σ 1′ + σ ′3 ) / 2 (= mean effective normal stress)


q = (σ 1′ − σ 3′ ) / 2 (= deviatoric stress/2)
TSP = total stress path

Figure 4.7: Effective stress paths for varying B values and total stress ratios (Folkes
and Crooks 1985).

Note: Ec = end of construction


Figure 4.8: Effective stress paths and pore water pressure generation associated with
different types of yield conditions (Leroueil and Tavenas 1986).

Muir Wood (1990) considered the undrained response of a soil element beneath the
centre of a circular load on a lightly over-consolidated, soft clay foundation (Figure
4.9). He described the soft clay using the Cam Clay model, assuming the soil to behave
isotropically and elastically prior to yield. A summary of the behaviour of the
foundation under the applied load as described by Muir Wood is as follows:
Chapter 4: Embankments on Soft Ground Page 4.12

• The total stress path is dependent on the total stress ratio and can be determined
from elastic analysis (Figure 4.10 for circular load).
• The effective stress changes (undrained response) are fixed by the stress-strain
behaviour of the soil and are largely independent of the total stress changes. Prior to
yield, the effective stress path shows no change in mean effective stress until the
yield locus is reached at Point C´ (Figure 4.11).
• Once the soil has yielded at Point C´, the effective stress path turns toward the
critical state at D´ and an increase in the rate of pore water pressure response occurs.
• The effective stress path B´C´D´ (Figure 4.11) is the undrained effective stress path
that is followed whatever the total stress path, provided it involves an increase in the
deviator stress, q.
• The total stress changes can be deduced from elastic analysis up to the point of yield
at C´. Once yielding has occurred the elastic analysis is no longer strictly valid
because the yielding elements of soil have a lower stiffness for continual increase in
shear stress, but it will provide a reasonable estimate so long as only a small
proportion of the total soil has reached yield.
• The initial excess pore water pressure response in undrained loading (Figure 4.12)
up to the point of yield arises entirely from the change in total mean stress that has
developed at that depth (Figure 4.10). Thus, for a soil element at depth z / a = 0.48 ,
∆u = 0.567 ∆ζ (where ∆ζ is the change in applied load) up to the point at yield at
C´.
• Once yielded, the excess pore response is made up of two parts; one due to the
applied change in total mean stress and the other due to the change in effective mean
stress as a consequence of yielding. The total stress path may or may not have been
affected by the occurrence of yield, and therefore it is not possible to make precise
statements about the excess pore water pressure response from C to D (Figure 4.12).
• Local failure is reached at Point D´ on the critical state line. At this point the soil is
unable to support any further stress, however, the local failure is contained and
supported by the surrounding unfailed soil. Assuming the soil is not strain
weakening, the effective stress path stops at Point D´, no further change in deviator
stress can occur and deformation continues without further change in effective
stress. This implies that the pore water pressure changes at Point D´ are equal to the
change in total mean stress and hence to the change in total vertical stress (Figure
4.10). The total stress path moves sideways at constant q (path DE in Figure 4.11).
Chapter 4: Embankments on Soft Ground Page 4.13

With respect to the pore water pressure response on loading, the idealistic example
given by Muir Wood (1990) summarised above assumes an undrained response of the
soft clay foundation. As discussed, the effective stress path up to the yield locus (path
B´C´ on Figure 4.11) is independent of the total stress path and the pore water pressure
response arises entirely from the change in total stress in undrained loading. Based on
field observation, Leroueil et al (1978a) indicated that the typical pore water pressure
response (Figure 4.3) observed during the initial stages of loading of an embankment on
soft ground (whilst the foundation is in an over-consolidated state) is significantly
different to that predicted by an undrained response because of consolidation and partial
drainage during loading whilst the foundation is in an over-consolidated condition.
Leroueil and Tavenas (1986) indicated that the amount of consolidation varies from
case to case, depending on the coefficient of consolidation of the over-consolidated
clay, drainage path length, and rate of loading. Jardine and Hight (1987) concur that the
field observation of pore water pressure response during the initial stages of loading
(whilst the foundation is in an over-consolidated condition) are indicative of
consolidation and therefore partial drainage. Folkes and Crooks (1986) comment that
consolidation due to partial drainage can occur during the initial stages of embankment
construction, but they consider that the incidence of near undrained conditions is
prevalent in a relatively large number of cases (approximately 35%).

Figure 4.9: Element of soil at depth z beneath on centreline of uniform circular load
(Muir Wood 1990)
Chapter 4: Embankments on Soft Ground Page 4.14

Note:
Total stresses are dimensionless.
α z = vertical stress
α r = horizontal stress
α p = mean stress
α q = deviator stress

Figure 4.10: Total stresses on centreline beneath uniform circular load on elastic half
space (Muir Wood 1990).

Note: p and p' are the mean total and mean effective stresses of the form (σ 1′ + 2σ ′3 ) / 3
q is the deviatoric stress defined as (σ 1′ − σ 3′ ) .

Figure 4.11: Total (BCDE) and effective (B′C′D′E′) stress paths for a soil element on
the centreline beneath a circular load (Muir Wood 1990).
Chapter 4: Embankments on Soft Ground Page 4.15

Note:
α p is the dimensionless total mean stress
defined as ∆p / ∆ζ
α z the dimensionless vertical stress defined as
∆σ v / ∆ζ
∆ζ is the change in applied load
Points BCDE relate to total stress states BCDE in
Figure 4.11

Figure 4.12: Variation of pore water pressure with applied surface load for element on
centreline beneath circular load (Muir Wood 1990).

In summary, the published literature on field case studies of embankments constructed


on lightly over-consolidated (OCR < 2.5) soft foundations indicates that in the majority
of cases partial drainage does occur in the initial stages of embankment construction
whilst the foundation is in an over-consolidated condition. However, in a number of
case studies the pore water pressure response of individual or a group of piezometers
during initial construction was close to that of an undrained response. Based on
published information and analysis of the 16 cases as part of this study (Appendix C,
Section 1) it is concluded that it is difficult to predict the pore water pressure response
during the over-consolidated stage of loading with any degree of accuracy. However, it
is important to closely monitor the excess pore water pressure response for assessment
of the effective stress state of the foundation for prediction of the likely deformation
behaviour. To do this it is necessary to monitor:
• The effective stress path, as recommended by Folkes and Crooks (1985), to assess at
what location it is likely to intersect the limit state surface (either below or above the
critical state line, Figure 4.7)
• The rate of pore water pressure generation ( B1 as recommended by Tavenas and
Leroueil (1980)) to assess the change in state of the foundation from an over-
consolidated to a normally consolidated condition.

4.2.2 DEFORMATION BEHAVIOUR

Marche and Chapuis (1974) analysed the horizontal displacements at the toe of eight
embankments as a function of the factor of safety during construction (Figure 4.13).
Chapter 4: Embankments on Soft Ground Page 4.16

They concluded that when the factor of safety falls below 1.4 the horizontal deformation
behaviour changes significantly indicating the approach of failure, and can therefore be
used to monitor the stability of embankments. Tavenas et al (1979) analysed the
horizontal deformation behaviour at the toe of some 21 embankments and concluded
that the change in deformation behaviour identified by Marche and Chapuis (1974) was
related to the change associated with going from a partially drained, over-consolidated
condition to a normally consolidated, undrained condition as previously discussed in
Section 4.2.1.
Tavenas et al (1979) proposed an empirical correlation, based on their case study
analysis, between the settlement under the centre of the embankment and the maximum
lateral displacement measured at the toe (from inclinometer observations). The
definition of parameters is shown in Figure 4.14 and the empirical correlation presented
in Figure 4.15. In the context of limit state theory, Tavenas et al (1979) and Tavenas
and Leroueil (1980) describe the deformation behaviour as follows:
• During the early stages of construction of an embankment on soft clay (whilst the
foundation is in an over-consolidated condition) deformations are small and partial
drainage is observed due to the relatively high modulus and high coefficient of
consolidation of the foundation. Lateral deformations at the toe are relatively small
in comparison to the settlement under the centre of the embankment due to the
effects of partial drainage, with y m ˜ 0.18s (standard deviation = 0.09), where ym is
the maximum lateral displacement at the toe and s is the settlement under the centre
of the embankment measured at any point in time (Figure 4.14 and Figure 4.15).
They indicated that the lateral displacement during this stage could probably be
predicted by the theory of elasticity adopting drained parameters for the clay
foundation and a Poisson’s ratio of about 0.3.
• A significant change in the deformation behaviour occurs once the effective stress
path intersects with the limit state surface (Point P´ on Figure 4.4) and the threshold
height, Hnc, is reached. As the threshold height is reached the behaviour of the
foundation changes from over-consolidated, partially drained conditions to normally
consolidated, undrained conditions. The effective stress path for increased loading
beyond Point P´ follows the path P´F´ of the now normally consolidated clay.
Undrained shear deformations occur under constant effective stress and the stress-
strain behaviour corresponds to plastic flow of the undrained normally consolidated
Chapter 4: Embankments on Soft Ground Page 4.17

clay. Deformations are much larger and both settlement and lateral displacements
increase at about the same rate ( ∆y m ≈ 0.91∆ s ) as shown on Figure 4.15.

Tavenas et al (1979) comment that the ratio ∆y m / ∆s is dependent on a variety of

parameters including the factor of safety and the angle of the embankment slope.
Plotting the ratio ∆y m / ∆s with time (Figure 4.15) is useful as a guide to the change

from over-consolidated conditions to normally consolidated, undrained conditions (i.e.


the onset of yielding of the foundation).

Notes:
∆q increase in vertical load producing
instantaneous horizontal displacement at
the surface ρ h
ρ h Eu
R=
∆ q.B

Figure 4.13: Variation of the dimensional factor R (proportional to the horizontal


surface displacement at the toe) with factor of safety (Marche and Chapuis 1974)
Chapter 4: Embankments on Soft Ground Page 4.18

Figure 4.14: Definition of the geometry and deformation parameters (Tavenas et al


1979)

Figure 4.15: Average correlation between the maximum lateral movement at the toe, ym,
and settlement under the centre of the embankment, s, during construction (Tavenas and
Leroueil 1980)

Tavenas et al (1979) also comment that the lateral displacement profile at the
embankment toe from inclinometer records is related to the thickness of the normally
consolidated clay layer at any stage during and post embankment construction (noting
that it increases as the load is applied). Figure 4.16 presents the observed effective
stress and lateral displacements profiles for a case where the full thickness of the soft
clay foundation is normally consolidated (Cubzac-A) and where only a part of the soft
Chapter 4: Embankments on Soft Ground Page 4.19

clay foundation is normally consolidated (St. Alban-B). For St. Alban-B, greater lateral
movements are observed within the zone of normally consolidated clay and are
associated with yielding (or plastic flow).

Note: the stresses ( σ ′vo , σ ′v , and σ ′p ) are defined in Figure 4.2


y is lateral displacement at the embankment toe
z is the depth below ground surface
Figure 4.16: Effective stress and lateral displacement profiles at the end of construction
under (a) Cubzac-les-Ponts A and (b) Saint Alban B test fills (Tavenas et al 1979)

Marsland and Powell (1977) considered that, for the Thames test embankment, the most
sensitive means of predicting the onset of failure was to monitor the ratio of the total
horizontal movement (at the toe of the embankment) to that occurring during actual
Chapter 4: Embankments on Soft Ground Page 4.20

application of the load. Figure 4.17 presents their data comparing the movement during
placement of one lift on the embankment (typical 5 hour placement time) with the
movement over a 24 hour period (including movement during loading) as the
embankment height is increased. This approach requires placement of lifts in uniform
layer thicknesses and in the case of Thames, at a rate of one lift per day. This method
incorporates the observations of; increase in lateral deformation as the effective stress
state moves from over-consolidated to normally consolidated (discussed above);
increase in lateral deformation as the zone of normally consolidated foundation
increases with increasing embankment height; and the theory of creep type movements
where the rate of creep type deformation increases with increasing levels of shear stress.

Figure 4.17: Comparison of horizontal movement (on berm beyond embankment toe)
during loading and for 24 hour period for the Thames test embankment (Marsland and
Powell 1977)

Ramalho-Ortigao et al (1983a) measured the horizontal displacements at the base of


the embankment of the Rio test embankment. The observed displacements (Figure
4.18) indicate that the greatest displacement occurs under the slope of the embankment.
This is possibly a reflection of the concentration of the shear strain at depth under the
Chapter 4: Embankments on Soft Ground Page 4.21

embankment slope; however, this effect could be masked by stiffness variations


between the soft clay and upper, weathered and heavily over-consolidated crust, and
between the crust and embankment.

Figure 4.18: Observations of lateral displacement at the base of the embankment with
increasing embankment thickness at the Rio test embankment (Ramalho-Ortigao et al
1983a)

Finite element modelling of numerous case studies of embankments on soft ground is


reported in the literature. Whilst the predicted results of deformation behaviour,
particularly lateral deformation, are in general relatively poor, what the results
collectively indicate is the development of high concentrations of shear stress at depth
below the slope of the embankment (Figure 4.19). Relatively high shear stress
concentrations are modelled at embankment thicknesses significantly below the
reported failure thickness; in the case of the Muar-F test embankment the stresses in
Figure 4.19 are at a fill thickness of 74% of the eventual failure thickness (Indraratna et
al 1992). Associated with this concentration of shear stress is a concentration of shear
strain. The implications of this concentration of shear stresses and strains in relation to
the observed lateral deformation profile in inclinometers at the toe of the embankment
are discussed in Section 4.3.
Chapter 4: Embankments on Soft Ground Page 4.22

Figure 4.19: Contours of normalised shear-stress ratios at 4 m fill thickness from finite
element modelling at Muar-F test embankment, Malaysia (Indraratna et al 1992)

4.3 ANALYSIS OF THE B EHAVIOUR OF F AILURE CASE S TUDIES

4.3.1 CASE STUDIES ANALYSED

The deformation behaviour and excess pore water pressure response of thirteen case
studies of well-monitored embankments on soft ground that were constructed to failure
have been analysed. Selection of suitable case studies was based on the level of
available information on the properties of the foundation and embankment filling, and
the amount, type, location and available published results of instrumentation. A
summary of each case study is given in Table 4.1.
Except for Mastemyr, which is a circular embankment, all the embankments are
rectangular in shape. In the case of Sackville the embankment is geotextile reinforced.
The embankment failure height at Sackville was assessed to have occurred when the
foundation failed, as the embankment did not fail until several additional metres of
filling had been placed. Two embankments have been included from the Muar test
embankments, Malaysia; Muar-F was purposely constructed to failure at a constant rate
of construction (Brand and Premchitt 1989), and Muar-EC was one of a series of
thirteen trial embankments constructed to assess various methods of ground
improvement (Malaysian Highway Authority 1989a). Prior to construction, the
foundation for the Muar-EC embankment was electro-chemically injected with Condor-
SS, a sulphonated petroleum based ion exchanger, to a depth of 7.6 m with hole
Chapter 4: Embankments on Soft Ground Page 4.23

spacings at 2 m centres. The main purpose of the Condor-SS chemical was to “reduce
the bound water in order to subsequently eliminate cavities and reduce the swell
capacity of the individual soil particles” (Malaysian Highway Authority 1989a).
Details of the Muar series of trial embankments are given in Appendix C (Section 2).
The foundations at the Kalix, St. Alban, James Bay, Portsmouth and Mastemyr
embankment sites are of sensitive clays with high undrained brittleness. For the
remainder of the case studies the foundations are of low sensitivity clays with limited
strain weakening in undrained loading.

In the context of identifying indicators of an impending failure condition, the


monitored observations analysed (Figure 4.20) are as follows:
• Excess pore water pressures under the centre and under the toe of the embankment;
• Settlement under the centre of the embankment (Point 1);
• Vertical deformation at the toe of the embankment (Point 2) and at a point beyond
the toe (Point 3), but within the zone of the eventual failure;
• Lateral deformation at the toe of the embankment (Point 2);
• Lateral deformation with depth (from inclinometer records) at the toe of the
embankment. Both the vertical inclination angle in the vicinity of the surface of
rupture (as an assessment of shear strain) and the maximum lateral deformation have
been analysed.

Embankment h Embankment
Height, H v
1 2 3
Weathered Crust

+ +
Soft Clay
Inclinometer
+ + plot

Inclinometer
+ +
LEGEND
+ Piezometers
Surface
Deformation point

Dense/Hard Foundation
Figure 4.20: Idealised section of embankment indicating the types of monitored
observations.
Chapter 4: Embankments on Soft Ground Page 4.24

Table 4.1: Summary of failure case studies of embankments on soft ground.


1
Material Properties * Analysis of Excess Pore Pressure Response
Failure Rate of
Slope
Name Height , Construction Foundation Threshold Height, B1 (refer Figure
(H to V) Embankment B2 (refer Figure 4.5)
H (m) (m/day) Crust Soft Clay Hnc 4.5)
Narbonne, France 9.6 3.7 to 1 1.9 Gravelly material 3m thick, CL clay, IL < 1 9m thick, CL/ML, soft, layered, IL = 1.1 - 1.5 Est ~ 5.5 m na na

Silty sand, loosely 8.8 m thick, CH clay, very soft, marine deposition, 1 m - (1.5 to 4.5 m 0.25 - 0.35 (1.5 to 4.5 0.8
Rio de Janiero, 2.4 m thick, CH clay, fissured, IL > 1
2.8 2 to 1 0.10 placed, some relatively homogeneous, low sensitivity, ductile, IL = 1.2 - depth) m depth) (1.25 to 2.25 for H >
Brazil below 1.2 m
cohesion 1.3 > 2.8 m - (from 4.5 m) 0.6 - 0.75 (> 4.5 m) 2.8 m)

Weathered
5.1 m CH clay, soft, tidal marine deposition, layered 0.85 (upper clay), 1.15
Kings Lynn, UK 6.7 1.4 to 1 0.67 sandstone, 1.2 m thick, CH silty clay, IL < 1 1.9 to 2.1 m 0.1 - 0.4
(peat layers), IL = 1.0 - 1.5 (Peat and lower clay)
compacted

4.1 (incl 1 16 m thick, MH/OH silt overlying CH clay, soft to very 1.4 to 1.7 at P2 on the
Crushed rock, soft, marine deposition, homogeneous, fissured, surface of rupture, B2 =
Kalix, Sweden m cut at 1.4 to 1 0.35 2 m thick, CH/MH silty clay, fissured 2.2 to 2.9 m 0.47 to 0.75
loosely placed sensitive, brittle, I L = 0.6 - 1.0
toe) 11 for H > 2.82 m

6.3 (incl 1 9 m thick, CL/ML, soft to very soft, tidal marine


Sandy gravel, loosely 3 m thick, CH organic silty clay, with
Thames, UK m cut at 2.6 to 1 0.10 deposition, layered (peat layers). ML/SC below 7.3 m, 1.0 to 1.6 m 0.18 - 0.4 0.8 to 1.0
placed 0.5 m layer of organic growth
toe) firm

12.2 m thick, CL/ML, soft to very soft, leached marine 2.0 m at depth of 1.5 m.
St. Alban-A, Uniform sand, 1.5 m thick, CL/CH clay, IL = 0.5 to 0.5 (base of crust) to
4.02 1.5 to 1 0.45 deposition (Champlain Sea Clay), cemented, sensitive, NC zone increases with 0.07 to 0.65
Quebec, Canada compacted 1.4 0.92
brittle, IL = 2 to 3 fill height.

Cubzac-A, France 4.5 1.5 to 1 0.26 - 0.75 cohesionless ? 1.8 m thick, CH clay, IL = 0.3 to 1.0 7.2 m thick, CH clay, soft to very soft, IL = 1 to 5.5 1.5 to 2.5 0.4 to 0.7 0.7 to 1.7

Granular material 5.5 m thick, 0.9 m organic CL/OL silt 9.5 m thick, CL/ML marine clay, soft, leached, very
James Bay, Canada 7.6 2 to 1 0.15 (sand to cobbles), overlying weathered CL clay, IL = 1.0 - sensitive, brittle, IL = 1.3 to 3, contains shells. 3.9 to 4.8 m 0.3 - 0.65 0.85 to 1.1
loosely compacted 1.1, fissured, very sensitive.

2.4 m thick, 0.3 m dense swamp 11 m thick, CL marine clay, soft to very soft, leached,
Portsmouth, USA, 2 to 3 in upper clay and
6.55 4 to 1 0.19 - 0.42 Clean sand overlying CL/CH silty clay, IL = 0.5 - sensitive, brittle, IL = 1.3 to 2.3. Contains occassional 0.45 - 0.65 0.95 - 1.0
New Hampshire SM/ML
1.3, fissured. shells. Underlain by SM/ML and soft clay.

Sand, some 2.5 m thick, MH/OH fibrous peat, IL = 15 m thick, CL/ML marine clay, very soft, leached, very
Mastemyr, Norway 2.84 1 to 1 0.19 sensitive (quick), brittle, fissured, IL = 2 to 4. 2 m (?) 0.75 - 1 1 - 1.7
compaction 0.35 - 1.0, very soft.
Gravelly silty sand 1.2 m thick, organic clayey silt (ML), 8.8 m thick, ML-MH clayey silt, soft to firm, intertidal salt
Sackville, Canada,
5.7 1 to 1 0.14 - 1.6 with some clay, matrix of fibrous material, IL = 0.6 - marsh deposit, layered, IL = 2 to 2.5 (to 6 m depth), OCR 1.5 to 2.4 m 0.6 - 0.75 1
New Brunswick
loosely placed 0.8. ~ 3.6 to 1.2.

Decomposed 16 m thick, CH marine clay. Upper clay layer (6 m thick)


2 m thick, weathered CH marine 1 m in upper clay 0.15 (upper clay)
Muar-F, Malaysia 5.4 2 to 1 0.054 Granite, CH/SC, very soft, lower clay layer soft, I L = 1 to 1.5, some shells, 1.0 - 1.2
clay, IL < 1, fissured. >3m in lower clay 0.8 (lower clay)
compacted ductile, low sensitivity

Muar-EC, Malaysia Decomposed 15 m thick, CH marine clay. Upper clay layer (6 m thick)
2 m thick, weathered CH marine
(Electro-chemical 4.68 2 to 1 0.017 Granite, CH/SC, very soft, lower clay layer soft, IL = 1.25 to 2, some 1.5 to 1.9 0.4 - 0.6 0.7 - 1.6
clay, IL < 1, fissured.
injection) compacted shells, ductile, low sensitivity
1
Note: * Soils classified to the Australian Soil Classificastion System (Australian Standard AS1726 - 1993: Geotechnical Site Investigations)
I L is the liquidity index = (wFMC - wP)/(wL - wP), where wFMC = field moisture content, wL = liquid limit and wP = plastic limit
na indicates data or information not available
Chapter 4: Embankments on Soft Ground Page 4.25

Table 4.1 (cont.): Summary of failure case studies of embankments on soft ground (Sheet 2 of 2).
Pre Failure Deformation
Geometry of
Lateral
Name Settlement Observations Leading up to Failure Post-Failure Deformation Surface of References
Displacement
(centre), mm Rupture
(toe), mm
Narbonne, France 700 na No cracking prior to failure. na Rotational Pilot (1972)
Embankment cracked at H = 2.5 m. Raised to 2.8 m Ramalho-Ortigao et al (1983a & 1983b)
Rio de Janiero, Suspect slow to moderate rate of movement at
295 155 and severe cracking but no failure. Failure during Rotational Costa-Filho et al (1977)
Brazil failure, displaced ~ 700 mm.
placement of lift above 3.1 m. Alameida & Ramalho-Ortigao (1982)
Embankment cracking observed on day before failure
Suspect possibly rapid rate of movement. Crest Complex - Rotational Wilkes (1972)
Kings Lynn, UK 460 > 125 (H = 6.15 m). Failure morning after construction to
settled ~ 1.25 m and toe heaved ~ 1.5 m. and Translational Wroth & Simpson (1972)
final height.

No embankment cracking prior to failure. Significant


Moderate rate of movement (max. rate ~1100 Holtz & Holm (1979)
Kalix, Sweden 755 > 105 heave and settlement on placement of last layer, Rotational
mm/day). Total movement at failure ~700 mm Stille et al (1976)
failure 3 hours later.

Possible embankment cracking prior to failure. Moderate rate of movement (max. rate ~700
Complex - Rotational
Thames, UK na 450 Appreciable movement on placement of final layers, mm/day). Total movement at failure ~2200 mm. Marsland & Powell (1977)
failure soon after completion of final layer. Full slip developed over 10 hours. and Translational

No embankment cracking prior to failure. At failure, Rapid rate of movement (~ 900 mm/min). Total
St. Alban-A,
150 > 40 most of movement occurred in 1 min followed by slow movement at failure ~ 1150 mm at crest and toe Rotational La Rochelle et al (1974)
Quebec, Canada
creep for next 10 mins. heave of 1200 mm.

Magnan et al (1982)
Cubzac-A, France 170 130 No indication of embankment cracking prior to failure. Total movement on surface of rupture ~ 1.4 m. Rotational Blondeau et al (1977)
Josseaume et al (1977)
No embankment cracking observed prior to failure.
Rapid failure at maximum rate of ~1800 mm/min.
Rise in water level in IN tube at toe observed prior to
James Bay, Canada 120 > 35 Total movement ~2.5 - 3 m crest settlement and Rotational Dascal & Tournier (1975)
failure. Failure occurred morning after construction to H
1.5 - 1.8 m toe heave.
= 7.6 m.

No embankment cracking observed prior to failure. Suspect likely rapid failure, but no observations.
Portsmouth, USA,
160 > 30 Failure occurred during the night, after construction to Total movement of 3 m crest settlement and similar Rotational Ladd (1972)
New Hampshire
7.6 m. toe heave.

Circular embankment so not plane strain conditions.


Mastemyr, Norway 700 215 na na Clausen (1972)
Failure occurred over-night.
Cracking observed in morning following construction to
Sackville, Canada, Rowe et al (1995)
520 > 410 5.7 m, also large increase in strain gauges on na Rotational
New Brunswick Rowe et al (1996)
geotextile and shearing of INs.

Raised to 5.4 m, cracking observed following morning Rapid failure (~3 mins) at estimated maximum rate Brand & Premchitt (1989)
370 (at 4.5 m
Muar-F, Malaysia 640 with slow increase in width over day. Next day raised to of ~1000 mm/min. Total movement ~ 1.4 m crest Rotational Poulos et al (1990)
depth)
5.6 m with failure on layer completion. settlement. Indraratna et al (1992)
Muar-EC, Malaysia Embankment cracking observed 11 days after Suspect moderate rate of failure. Initial 340 mm
(Electro-chemical 720 200 construction to 4.68 m. Further cracking developed movement followed by slow continued movement Rotational Malaysian Highway Authority (1989a)
injection) with failure 31 days after construction. for many days.
Chapter 4: Embankments on soft ground Page 4.26

4.3.2 EXCESS PORE WATER PRESSURE

The observed excess pore water pressures in selected piezometers under the centre of
the embankment during embankment construction (of the analysed case studies) are
presented in Figure 4.21, and includes the excess pore water pressure response for
B = 1 for comparative purposes. The critical effective vertical stress ( σ 1′ .crit ),
describing the change from an over-consolidated to a normally consolidated condition,
assessed from the excess pore water pressure response, is indicated on Figure 4.21 and
is summarised in Table 4.2 for the piezometers plotted.

Table 4.2: Excess pore water pressure observations for the failure case studies
Name ′
Piezometer σ vo Rate of B1 B2 σ 1′.crit σ ′p Hnc
Construction (lab)
(depth) (kPa) (m/day) (kPa) (kPa) (m)

Muar-F P2 (5 m) 27 0.054 0.23 1.0 44 44 1.1


King’s Lynn P215 (4.7 m) 37 0.67 0.1 1.15 68 - 1.9
1
Muar-EC P4 (4.6 m) 31 0.017* 0.37 to 0.5 1.0 to 1.5 51 44 1.5
Portsmouth V5 (5.8 m) 55 0.19 – 0.42 0.6 0.98 62 69 2.0
Thames C1 (2.3 m) 11.6 0.10 0.18 0.98 27 24 1.0
Sackville P15 (2 m) 18 0.14 – 1.6 0.66 0.97 28 43 1.5
James Bay A3 (7.6 m) 62 0.15 0.66 1.1 88 125 3.9
Kalix P1 (2.5 m) 17 0.35 0.4 to 0.8 1.0 to 1.7 35 38 2.2
P5 (5 m) 24 0.4 to 0.65 1.2 to 11 36 42 2.4
Rio P2B (3 m) 9.6 0.10 0.27 0.80 23 22.5 1.0
Mastemyr P1 (3.5 m) 16 0.19 0.5 to 1.0 1.0 to 2.0 20 24 1.3
St. Alban-A C1 (3 m) 26 0.45 0.2 to 0.4 0.92 39 45 1.7
Cubzac-A P112 (4 m) 32 0.26 – 0.75 0.4 1.0 to 1.7 53 51 1.8
1
Nore: * Muar-EC was constructed in stages at an overall average rate of 0.017 m/day.
′ = initial vertical effective stress (at the piezometer)
σ vo
B1 and B2 are defined in Figure 4.5
σ 1′.crit = effective vertical stress at the threshold height (at the piezometer)
σ ′p = pre-consolidation pressure (at the piezometer)
Hnc = threshold height at the piezometer (defined in Equation 4.5)
Chapter 4: Embankments on soft ground Page 4.27

120
a)
Kalix - P1 (2.5 m)
Rio - P2B (3 m)
100
Mastemyr - P1 (3.5 m)
Excess Pore Water Pressure (kPa) .

Kalix - P5 (5 m, crest)
St. Alban A - C1 (3 m)
80
Cubzac-A - P112 (4 m)
B =1

60 Effective Critical Stress

40

20

0
0 10 20 30 40 50 60 70 80 90 100

Applied Embankment Load (kPa)

120
b)
Muar F - P2 (5 m)
Kings Lynn - P215 (4.7 m)
100 Muar EC - P4 (4.6 m)
Portsmouth - V5 (5.8 m)
Excess Pore Water Pressure (kPa) .

Thames - C1 (2.3 m)
80 Sackville - P15 (2 m)
James Bay - A3 (7.6 m)
B =1

60 Effective Critical Stress

40

20

0
0 20 40 60 80 100 120

Applied Embankment Load (kPa)

Figure 4.21: (a) and (b) Excess pore water pressure under centre of embankment versus
applied embankment load.

In general, the excess pore water pressure response during embankment construction is
in accordance with the limit state interpretation by Leroueil et al (1978a) with an initial
partially drained response followed by an undrained response ( B 2 ≈ 1 ) as discussed in
Section 4.2.1 and shown on Figure 4.5. Only Piezometer P5 at Kalix (Figure 4.21a) is
considered to represent the pore water pressure response at the onset of failure; i.e.
Chapter 4: Embankments on soft ground Page 4.28

B f > 1 (Figure 4.5). This is because the piezometer is considered to be located close to

the eventual surface of rupture and within the initial localised failure zone. At an
applied load of 64 kPa the total pore water pressure at Piezometer P5 equates to
approximately the total effective vertical stress and is considered to represent static
liquefaction on the surface of rupture prior to the eventual embankment failure.
From Table 4.2 it is evident that the estimated effective critical stress (σ 1′, crit ) from

the excess pore water pressure response, in most cases, reasonably approximates the
laboratory estimated pre-consolidation pressure, σ ′p , taking into consideration the

accuracy of the estimation of the critical stress, sampling disturbance and accuracy of
the laboratory procedures. Therefore, the excess pore water pressure response under the
embankment centreline is a good indicator of the change from partially drained over-
consolidated to undrained normally consolidated conditions for lightly over-
consolidated clay foundations (OCR < 2.5). The exception, Piezometer A3 at James
Bay (refer Table 4.3), is discussed further in Appendix C (Section 3).
The values of B1 are quite variable as indicated in Table 4.2. Leroueil and Tavenas
(1986) commented that the degree of consolidation during the initial construction period
(whilst the foundation is in an over-consolidated condition) varies from case to case,
depending on the coefficient of consolidation, cv, of the over-consolidated clay, the
length of the drainage path and the rate of loading. Crooks (1987) considers that the
clay stiffness and permeability, rate of loading and drainage boundary conditions affect
the degree of consolidation during initial loading. He comments that clay foundations
with high stiffness (in an over-consolidated condition) would be expected to generate
only moderate excess pore water pressure response during initial loading whilst clays of
lower stiffness would be expected to develop significant excess pore water pressure
response. However, most of these assertions (except for drainage path length, Leroueil
et al (1978a)) are intuitive and little case study evidence is given in support.
With respect to rate of construction, Leroueil et al (1978b) provide field evidence
from the St. Alban test site that shows the initial rate of pore water pressure generation
( B1 ) for embankment A (constructed at approximately twice the rate of other
embankments) was greater than that for embankments B, C and D (Appendix C, Section
1). Leroueil et al (1978b) attributed this to the reduced consolidation associated with a
faster rate of construction.
Chapter 4: Embankments on soft ground Page 4.29

To evaluate the influence of the properties of the foundation and rate of loading on
the rate of pore water pressure generation during the initial stages of construction prior
to yielding of the foundation, the field observations from 16 case studies (54
piezometers) were analysed. The excess pore water pressure profiles from each case
were compared against each other as well as to an undrained response assuming an
infinite strip load on elastic half space. The results of the analysis are given in
Appendix C (Section 1) and the findings are summarised as follows:
• Virtually all cases indicate significant drainage occurs in the upper soil profile. The
excess pore water pressure profile (in comparison to the undrained profile for an
infinite strip on elastic half space) for most case studies is typical of a consolidation
isochrone for upward drainage only. Several case studies (Portsmouth, James Bay
and St. Alban A and B) are typical of a consolidation isochrone for drainage
boundaries at the surface and at depth. All four of these case studies have
foundations comprising low plasticity, sensitive, leached marine clays.
• There is a high degree of variability in the pore water pressure profile from case to
case with no clear correlation evident with respect to rate of loading (between
different sites), material type (sensitive, low plasticity clays compared to high
plasticity clays of low sensitivity) or likely material stiffness. For sites analysed
with more than one test embankment, the St. Alban test embankments (Leroueil et al
1978a) do indicate a correlation with respect to rate of loading as discussed above,
however, no correlation is evident between the Muar test embankments.
• Permeability is considered to being a key factor in controlling the variation of
coefficient of consolidation, cv, between case studies at different sites. In the case of
the Muar test embankments a significant variation in B1 was observed between the

upper and lower clay layers, with lower B1 values obtained in the more permeable

upper clay and much higher B1 values in the less permeable lower clay. A similar
pore water pressure profile to the Muar embankments was observed at the Rio test
embankment.
• In 7 of the 16 case studies analysed the rate of pore water pressure generation at a
depth below 4 to 6 m exceeded 80% of the pore water pressure profile assuming
undrained conditions. This indicates that near undrained conditions are observed at
depth in a significant proportion of cases during the initial stages of construction
whilst the foundation is in an over-consolidated condition.
Chapter 4: Embankments on soft ground Page 4.30

• The Mastemyr test embankment stands out as having a very high rate of pore water
pressure generation during the initial stages of construction, greater than that
estimated assuming undrained conditions. This observation may be due to the
assumptions of the foundation as a half space and that the foundation behaves
elastically prior to yielding for the undrained case considered.

For the piezometric response during normally consolidated conditions the analysed
case studies (Figure 4.21) indicate most of the excess pore water pressure observations
maintain a B2 of approximately 1, with several exceptions (refer Table 4.3). As
previously discussed, only in the case of Kalix is it considered that the excess pore
water pressure response is indicative of the onset to failure (i.e. B f > 1 , Figure 4.5).

The reason for the observation of constant B 2 leading up to a failure condition for most
piezometers is that the piezometers are not located within the localised failure zone. To
observe an onset to failure condition ( B f > 1 ), as is the case for Kalix, the piezometer

must be located within the localised failure zone of a strain weakening soil. In most
cases the zone of localised failure is relatively small, and coincident with the eventual
surface of rupture, and therefore it would generally be unlikely for a piezometer to be
located in this zone. Elsewhere in the foundation (away from the developed localised
failure zone), at the time of failure the effective stress conditions have not yet reached a
localised failure condition (Point F´ on Figure 4.4), and no change is therefore observed
for B 2 up to failure.
It is concluded that whilst the excess pore water pressure response is a good
indicator of the onset of undrained, normally consolidated conditions, it is not a reliable
indicator of an impending failure condition. This is because only piezometers located
within the relatively thin localised failure zone of strain weakening soils will the onset
of failure be observed (i.e. B f > 1).

It is important to note that the analysis of excess pore water pressure is generally for
piezometers under the centre of the embankment for which it can be assumed that the
equations for excess pore water pressure development apply (Equations 4.2, 4.3 and
4.4). The more general solution (proposed by Henkel (1961)) applicable to all stress
states is of the form:
Chapter 4: Embankments on soft ground Page 4.31

∆u = β∆σ oct + 3α∆τ oct (4.5)

where ∆σ oct is the octahedral normal stress, ∆τ oct the octahedral shear stress, and a
and ß are coefficients related to the soil behaviour.
The response of excess pore water pressure with time during periods of no
construction is also an important aspect in the assessment of stability and deformation
behaviour of embankments on soft ground. The excess pore water pressure response in
some piezometers increased during periods on no construction leading up to the failure
at Muar-EC, St. Alban-A and King’s Lynn embankments. A delayed failure was also
reported at Kalix, James Bay, Portsmouth and Mastemyr, where the failure occurred
less than 24 hours after the final fill layer had been placed. The observations and
mechanisms associated with increasing pore water pressure response during no
construction periods are discussed in Section 4.4.2.

Table 4.3: Summary of exceptions to general observations of excess pore water pressure
response.
Case Study Exception Comments / Reasons *1
James Bay σ ′p > σ 1′, crit (Table 4.2) Effective critical stress significantly lower than the
pre-consolidation pressure determined by laboratory
testing. This is thought to be due the to relatively
high over-consolidation ratio for the foundation
(OCR > 2.5) and effective stress path for this
condition.
Mastemyr B2 >> 1 (Table 4.2) Effective stress path possibly as per path OYE c in
Figure 4.8a. High B2 observed in piezometers down
to 6 m depth. Possibly a result of strain weakening
and indicative of partial breakdown in the soil
structure.
Kalix B2 >> 1 (Table 4.2) Effective stress path possibly as per path OYE c in
Figure 4.8a. High B2 observed in piezometers down
to 7.5 m depth. Possibly a result of strain weakening
and indicative of partial breakdown in the soil
structure.
Muar-EC B2 >> 1 (Table 4.2) Only Piezometer P4 gave high B2 . Uncertain as to
the cause of this observation.
Cubzac-A B2 >> 1 (Table 4.2) Piezometer P112 shows B2 in excess of 1.0. Only
limited information is available and it is uncertain as
to the cause of the observation.
Note: *1 The exceptions are discussed further in Appendix C (Section 3).
Chapter 4: Embankments on soft ground Page 4.32

4.3.3 PRE-FAILURE DEFORMATION BEHAVIOUR AND THE EFFECTS OF PROGRESSIVE


FAILURE

This sub-section discusses the pre-failure deformation behaviour of embankments on


soft ground leading up to a failure condition, and includes consideration of the
mechanism/s of failure, such as progressive failure.
As discussed in Section 4.2.2, the deformation behaviour of embankments on soft
ground is significantly affected by the effective stress state of the foundation in relation
to its limit state. At effective stress states within the limit state (i.e. an over-
consolidated condition) deformations will be small and partial drainage is observed due
to the low compressibility of the over-consolidated clay (and resultant high coefficient
of consolidation, cv) and small amount of pore fluid that flows from the soil. On
reaching the limit state surface (normally consolidated condition) deformations are
much larger due to yielding of the foundation and negligible pore water pressure
dissipation occurs (i.e. undrained conditions) due to the much lower cv value associated
with greater compressibility of the foundation.
Discussing the deformation behaviour of embankments on soft ground with respect
to the type of movement for brittle, sensitive clays as opposed to ductile clays Ladd
(1991) commented that:
• Brittle, sensitive soils show small undrained shear deformations during loading with
little warning prior to a well-defined failure. Post-failure the displacements are very
abrupt and large due to large strength reduction along a distinct surface of rupture,
and large pore water pressure increases occur along the rupture surface.
• Ductile soils show large undrained shear deformations during loading, especially at
low factors of safety, and provide ample warning prior to an ill-defined failure. At
and post failure, deformations are relatively small and occur within a zone rather
than along a distinct surface of rupture, with small pore water pressure changes.

Numerous authors (Ladd 1991; Rowe et al 1995; Marsland and Powell 1977; Fell et al
1987; Marche and Chapuis 1974) consider the horizontal movement at the toe of the
embankment to be a good indicator of an impending failure condition. Ladd (1991)
remarked that for ductile, soft clay foundations of low sensitivity the horizontal
movement at the toe provides the clearest evidence of an impending failure condition
because it is less affected by consolidation than settlement data and therefore reflects
more clearly the deformation caused by undrained shear.
Chapter 4: Embankments on soft ground Page 4.33

In the following sub-sections the deformation behaviour of the case study analysis
of failures of embankments on soft ground is discussed in the context of the pre-failure
deformation behaviour.

4.3.3.1 Vertical Displacement under the Centre of the Embankment


The settlement under the centre of the embankment versus relative embankment height
(H/Hf, where H is the embankment height and Hf is the embankment failure height) of
the analysed case studies is shown in Figure 4.22. The general behaviour, as expected,
indicates an increase in the rate of the settlement (with relative embankment height) at
embankment heights exceeding the threshold height, Hnc.
From the results it is considered that settlement under the centre of the embankment
does not provide an indication of an impending failure condition. The reason for this is
considered to be that in most cases the settlement monitoring point under the centre of
the embankment is not located within the failure zone.

4.3.3.2 Vertical Deformation At and Beyond the Embankment Toe

The vertical deformation at the toe of the embankment versus relative embankment
height for the analysed case studies is shown in Figure 4.23. Figure 4.24 presents the
vertical deformation behaviour beyond the embankment toe (approximately 5 m
distance beyond the toe). In all cases the monitoring point beyond the toe was located
within the eventual failure zone.
It is considered that the vertical displacement at and beyond the toe of the
embankment is a good indicator of an impending failure condition, particularly for
monitoring points located beyond the embankment toe (and within the eventual failure
zone). A measurable change in rate or direction of vertical displacement is observable
in most of the case studies beyond the threshold height and within the range 70 to 95%
of the eventual failure height.
For the monitoring points beyond the embankment toe (Figure 4.24), in most cases
negligible vertical deformations were observed during the initial period of embankment
construction, and the impending failure condition was identifiable by heave movements
or large increases in the rate of heave movement with increasing embankment height.
The observations therefore indicate that deformations at (and more significantly
Chapter 4: Embankments on soft ground Page 4.34

beyond) the toe are not significantly affected by the immediate settlement associated
with the embankment load, but are indicative of movements associated with instability
of the embankment.

0
a)

50

100
Rio - S2
Settlement (mm) .

Mastemyr - S1
150
Portsmouth - S1
James Bay - T12
200
St. Alban-A - R23

Threshold Fill Height


250

300

350
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

b)
0

200

400 Kalix - S2
Settlement (mm) .

Rio - S2
Muar-F - S1
600 Kings Lynn
Muar-EC
Sackville - 8S
800
Cubzac A
Narbonne
1000 Threshold Fill Height

1200
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

Figure 4.22: Settlement under the centre of the embankment versus relative
embankment height (H/Hf).
Chapter 4: Embankments on soft ground Page 4.35

-500
a)
Kalix - S4
-400 Narbonne - 5
Sackville - 4H (toe + 1m)
.
Vertical Displacement (mm),

-300 Threshold Fill Height


heave is neagative .

-200

-100

100
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

b) -80
Rio - S4
Muar-F - H9
-60 Mastemyr - PL17
James Bay - SMP 68
St. Alban-A - R13
Vertical Displacement (mm) .

-40
heave is neagative .

Threshold Fill Height

-20

20

40
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

Figure 4.23: Vertical displacement at the embankment toe versus relative embankment
height.
Chapter 4: Embankments on soft ground Page 4.36

-400
Sackville - 3H (toe + 2.5m)
-350 Rio - S6 (toe + 4m)
Muar - H7 (toe + 5m)
Narbonne (toe + 10m)
Vertical Displacement (mm) .

-300 Mastemyr - PL3 (toe + 3.9m)


Mastemyr - S13 (toe + 5.8m)
Heave is negative .

St. Alban - R8 (toe + 4.7m)


-250
Threshold Fill Height
-200

-150

-100

-50

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

Figure 4.24: Vertical displacement beyond the embankment toe versus relative
embankment height.

The cases where the vertical deformation at and beyond the toe is a good indicator of an
impending failure condition cover a wide variety of soil types from high plasticity clays
of low sensitivity to highly sensitive, low plasticity clays and silts, making it a good
indicator for embankments on all types of soft ground conditions.
For the embankments on highly sensitive, low plasticity clay foundations (St. Alban
and James Bay) the amount of vertical movement is relatively small (up to 10 to 15
mm) for up to approximately 90% of the eventual embankment failure height.
However, in recognising this, the vertical deformation at and beyond the toe is still
considered a good indicator for these types of foundation conditions as the purpose of
the monitoring would be to identify these small but measurable heave movements.
The vertical displacement at and beyond the toe of the Kalix and Rio test
embankments is considered as a reasonable to good indicator of an impending failure
condition, however, the deformation behaviour is not explainable only by being close to
the failure height (refer Table 4.4). For St. Alban-A the vertical deformation behaviour
at the toe is considered to be a reasonable indicator of an impending failure condition.
The heave movements start at about 60% of the failure height, however, the movements
are small (less than 10 mm up to 90% fill height) and a change in rate in not clearly
perceptible over the later stages of construction.
Chapter 4: Embankments on soft ground Page 4.37

4.3.3.3 Lateral Surface Displacements at the Embankment Toe


The lateral displacement at the toe of the embankment for the failure case studies has
been presented as the lateral surface displacement versus relative embankment height
(Figure 4.25).
From Figure 4.25 it is concluded that:
• In most cases an increase in the rate of lateral displacement relative to the
embankment height occurs at about the threshold height, Hnc, determined from the
pore water pressure response, which is in agreement with the interpretation of the
behaviour of embankments on soft ground to limit state theory.
• A further increase in the rate of lateral surface displacement occurs in most of case
studies analysed at relative embankment heights of 70 to 90%. It is considered that
this behaviour is an indication of an impending failure condition.
• For 11 of the 12 cases analysed, the lateral surface displacement is a good indicator
of impending failure. These cases cover a broad range of soil types from low to
high sensitivity and low to high plasticity, making it a good indicator for
embankments on all types of soft ground conditions.

The lateral surface displacement versus embankment height plots for Thames
(increase in rate beyond 75 to 80%), Muar-F (increase in rate beyond 75 to 80%),
Mastemyr (beyond 80%), Cubzac A (beyond 80%), Muar-EC (beyond 85%), St. Alban-
A (beyond 75%) and Portsmouth (beyond 80 to 85%) all show an increase in the rate of
displacement beyond the threshold height. For James Bay, Kalix and Sackville there
are insufficient data points beyond the threshold height to draw any firm conclusions,
however, the trend of the plots for these cases indicate the lateral surface displacement
at the toe is a good indicator.
For Kalix and Rio the change in deformation behaviour is not explainable only by
being close to the failure height, and at King’s Lynn the impending failure is not clearly
identifiable (refer Table 4.4).
The results indicate that the rate of embankment construction affects the amount of
lateral surface displacement measured, particularly for the case studies with high
plasticity, low sensitivity foundations. For embankment constructed at relatively slow
rates (Muar-F and Muar-EC constructed at rates of 0.02 to 0.055 m/day) the lateral
surface displacement was a good indicator of the impending failure, and for the
embankments constructed at more rapid rates (Rio at 0.1 m/day and King’s Lynn at 0.67
Chapter 4: Embankments on soft ground Page 4.38

m/day) the lateral surface displacement was not such a good indicator of the impending
failure. For Muar-EC it was evident that a significant proportion of the lateral
displacement occurred during rest periods in the later stages of construction, whilst the
rate during construction periods did not increase significantly. During these rest periods
at Muar-EC when the embankment height was above 50% of the eventual failure height,
the excess pore water pressure was either steady or else increased.

a) 500
Thames
Rio
Muar-F (at 4.5 m depth)
400 Kings Lynn
Mastemyr - PL17
.

Cubzac-A
Sackville
Lateral Displacement (mm)

Muar-EC
300 Kalix

Threshold Fill Height

200

100

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

b) 45

40 St. Alban-A - R6
James Bay
35 Portsmouth - SI-1
.
Lateral Displacement (mm)

30 Threshold Fill Height

25

20

15

10

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf
Figure 4.25: Lateral surface displacement at the embankment toe versus relative
embankment height.
Chapter 4: Embankments on soft ground Page 4.39

4.3.3.4 Lateral Displacement with Depth at the Embankment Toe


The data on lateral displacement with depth at the embankment toe, from inclinometer
observations, is presented in the form of the vertical inclination angle at the eventual
surface of rupture versus relative embankment height (Figure 4.27). It is recognised
that during construction the eventual surface of rupture is not known. However, by
monitoring the incremental vertical inclination angle (Figure 4.26) the zone of localised
failure, particularly for foundations of low sensitivity, will become evident prior to the
eventual failure. Once the localised failure zone becomes evident then the data from
this zone can be plotted in the form of Figure 4.27.
Figure 4.26 clearly shows the development of the localised failure zone with
increasing fill height and time for the Thames and King’s Lynn case studies. Ramalho-
Ortigao et al (1983a, 1983b) presented similar type plots for the Rio test embankment.
From Figure 4.26 it is evident that the eventual surface of rupture is located within the
localised failure zone.

a) 0 b) 0
Fill Hgt = 2.3 m Fill Hgt = 2.7 m
Fill Hgt = 3.3 m Fill Hgt = 3.8 m
2 Fill Hgt = 3.97 m 1
Fill Hgt = 4.4 m
Fill Hgt = 4.63 m
Fill Hgt = 6.1 m
Fill Hgt = 4.63 m
4 Fill Hgt = 6.1 m
Fill Hgt = 4.63 m 2
Fill Hgt = 4.97 m Fill Hgt = 6.7 m
6 Fill Hgt = 5.3 m Fill Hgt = 6.7 m
3
Surface of
Depth (m)

Rupture
Depth (m)

8
4

10 Surface of
5 Rupture

12

6
14

7
16

8
18
-5 0 5 10 15
-2 0 2 4 6 8 10 12
2 2
Vertical Inclination Angle (rad x 10 ) Vertical Inclination Angle (rad x 10 )

Figure 4.26: Incremental vertical inclination angle (from inclinometer observations at


the embankment toe) with fill height (and time) for (a) Thames and (b) King’s Lynn.
Chapter 4: Embankments on soft ground Page 4.40

25
a) Thames - IC (crest)
Rio
Muar-F
.

20 King's Lynn
Vertical Inclination Angle on Surface
of Rupture (radians x 100) .

Muar-EC
Sackville - I22 (toe)
Thames - ID (toe)
15 Mastemyr - I3

Threshold Fill Height

10

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Relative Embankment Height, H/H f

b) 3
Portsmouth - sensitive clay
Kalix
2.5 James Bay - ~8 m
.

Portsmouth - silt/sand
Vertical Inclination Angle on Surface
of Rupture (radians x 100) .

Threshold Fill Height


2

1.5

0.5

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

Figure 4.27: Vertical inclination angle on the surface of rupture from inclinometer
observations at the embankment toe versus relative embankment height.

Figure 4.27a is of predominantly low sensitivity clays (except for Mastemyr) and shows
that in all cases (except Rio and to a lesser extent King’s Lynn) significant increases in
the vertical inclination angle on the eventual surface of rupture occur at embankment
heights of 70 to 90% of the failure height. The increases in shear strain within the
localised failure zone and on the eventual surface of rupture occur well beyond the
Chapter 4: Embankments on soft ground Page 4.41

threshold height (Figure 4.27a) and are considered to represent a localised failure
condition within which large shear strains as a result of plastic deformation develop
(Figure 4.26). At the onset of localised failure the effective stress state within the
localised failure zone is considered to be at or close to the point of failure on the
effective stress path (Point F´ in Figure 4.4).
For embankments constructed on sensitive, brittle foundations (Portsmouth, Kalix
and James Bay in Figure 4.27b and Mastemyr in Figure 4.27a) the plot of maximum
vertical inclination angle versus relative embankment height as an indicator of an
impending failure is considered to vary from poor to good. Although the plots from
Figure 4.27 would indicate that it is a reasonable indicator of an impending failure
condition for these soil types, there are other explanations for the observed behaviour.
The development of the localised failure zone is considered to initiate within the
highly stressed zone of the foundation located under the shoulder of embankment
shoulder (as discussed in Section 4.2.2). At this point the soil in the localised failure
zone is unable to support any further stress, however, the local failure is contained and
supported by the surrounding unfailed soil (Muir Wood 1990). This is the start of a
progressive failure mechanism. With additional loading the size of the localised failure
zone increases and with increased straining the strength within the localised failure zone
decreases due to strain weakening. At some point the driving forces exceed the
resisting forces and failure occurs, and the mobilised strength at failure for strain
weakening soils will be less than peak strength.
It would be expected, and the results indicate, that the development of the localised
failure zone is more clearly developed for foundations of low sensitivity due to their
significantly lower undrained brittleness index (or lower amount of post peak strain
weakening). Once a localised failure zone has developed the amount of load transferred
to the surrounding soil is much less for the low sensitivity soils, due to their relatively
low undrained brittleness. In addition, relatively large post-peak strains are required for
strain weakening, and therefore a progressive failure mechanism develops relatively
slowly.
In the case of sensitive clay foundations, particularly those of low plasticity, a
progressive failure mechanism also occurs; however, it develops much more quickly
and possibly not until the embankment has been constructed to its failure height. For
these soils, peak strengths are reached at relatively small strains and at failure (of the
soil) a large reduction in the undrained strength occurs with relatively small strain. The
Chapter 4: Embankments on soft ground Page 4.42

large loss in undrained strength of these sensitive soils is associated with contraction of
the soil structure on shearing and transfer of load onto the pore fluid resulting in the
development of high excess pore water pressures; i.e. a static liquefaction condition.
Therefore, once a failure condition is reached progressive failure develops rapidly and
failure of the embankment quickly ensues. The observation of the deformation
behaviour within the inclinometer at the toe of the embankment at Portsmouth and
James Bay tend to confirm this rapid progression to failure.
For Portsmouth no significant increase in vertical inclination angle is evident in the
region of the eventual surface of rupture up to 98% of the embankment failure height
(Figure 4.27b). This indicates that the localised failure zone at Portsmouth is not likely
to have developed until some time after the end of construction. The embankment
failure occurred at some stage during the night (Ladd 1972), less than 12 to 18 hours
after construction to the eventual failure height. Within the sandy silt layer below about
10.4 m depth an increase in the vertical inclination angle (VIA) is evident from 68% of
the embankment failure height (Figure 4.27b).
The description leading up to failure of the James Bay embankment (Dascal and
Tournier 1975) strongly supports the rapid development of progressive failure as
discussed above. They state that:

“The failure started during the inclinometer readings by the field crew. At the
thirteenth measurement, they started to notice that that the order of magnitude
of the readings was quite different from the previous ones. At the fifteenth
measurement, the variation was such that the reading scale had to be changed.
At the twentieth measurement the probe became stuck and could neither be
lowered nor extracted. At the same moment, water started to rise in the tube.
Simultaneously they observed the rising of the natural ground at the toe of the
slope by about 0.3 m in approximately 10 sec, during which the fill was
ruptured.”

These observations would indicate the rapid development of shearing in the region
of the surface of rupture leading up to the failure and that significant shearing did not
take place until shortly prior to the failure. The observation of water rise in the
inclinometer tube would be consistent with static liquefaction on the surface of rupture
due to collapse of clay structure followed by the rapid failure of the embankment.
In the case of Kalix and Mastemyr it is considered that the partial breakdown in the
soil structure has reduced the undrained brittleness of the foundation and therefore
Chapter 4: Embankments on soft ground Page 4.43

progressive failure is more clearly evident than for the other case studies with sensitive
clay foundations.

4.3.3.5 Settlement Versus Lateral Displacement

Figure 4.28 presents the plot of settlement under the centre of the embankment versus
maximum lateral displacement at the embankment toe for ten of the thirteen failure case
studies analysed. The results indicate that an increase in the ratio of settlement to
maximum lateral displacement (∆y m / ∆s ) occurs above the threshold height and that

the ratio above the threshold height is quite variable. For those embankments that
cracked prior to failure (Rio, King’s Lynn, Muar-F and Muar-EC) an increase in the
∆y m / ∆s ratio occurs after cracking. No indication of the onset of failure is apparent
from the plot nor was it expected. Tavenas et al (1979) and Tavenas and Leroueil
(1980) used this type of plot as confirmation of the change in deformation behaviour
that occurs at the change from over-consolidated to normally consolidated conditions
and as field evidence to indicate that partial drainage does occur whilst the foundation is
in an over-consolidated condition.

450
Kalix
Rio
400 Muar-EC
Portsmouth
Mastemyr
350
Maximum Lateral Displacement .

Cubzac-A
Muar-F Embankment
Sackville cracked
300
at Toe, y m (mm) .

James Bay
Kings Lynn
250
Threshold Fill Height

200

150

100

Embankment
50 cracked

0
0 100 200 300 400 500 600 700 800

Settlement under Centre of Embankment, s (mm)

Figure 4.28: Maximum lateral displacement at the embankment toe versus settlement
under the centre of the embankment.
Chapter 4: Embankments on soft ground Page 4.44

4.3.3.6 Case Study Exceptions to Deformation as an Indicator of Impending Failure


The case studies for which deformation was a poor indicator of an impending failure
condition or the change in deformation behaviour has alternative explanations are
summarised in Table 4.4. Only those deformation indicators that provided a good
indication of an impending failure condition for most case studies have been considered;
vertical deformation at and beyond the embankment toe, lateral surface displacement at
the embankment toe and lateral displacement at depth at the embankment toe.

Table 4.4: Summary of case studies that are exceptions or have other explanations to
deformation as an indicator of impending failure.
Case Study Indicator Comments / Reasons *1
Vertical deformation at and Reasonable to good indicators.
beyond the toe (Section 4.3.3.2); Alternative explanation is that the large
lateral surface displacement at the increase in rate is associated with
embankment toe (Section 4.3.3.3) cracking of embankment.
Lateral displacement at depth at Poor indicator. Localised failure zone
Rio
the embankment toe (Section may not have developed until the
4.3.3.4) embankment had cracked. Possibly
influenced by the small embankments
constructed at the toe of the main
embankment.
St. Alban-A Vertical deformation at and Reasonable indicator, however, heave
beyond the toe (Section 4.3.3.2) movements are small (< 10 mm) and a
change in rate is not clearly perceptible.
Kalix All indicators *2 Good indicator, but alternative
explanation is that the change in
deformation behaviour is coincident with
the threshold fill height. Also, limitation
on number of measurement points taken
beyond the threshold height.
King’s Lynn Lateral surface displacement at Not a good indicator. An impending
the embankment toe (Section failure condition is not clearly
4.3.3.3) identifiable
Low Lateral displacement at depth at Incorporates the Portsmouth, James Bay
plasticity, the embankment toe (Section and St. Alban-A test embankments. Poor
sensitive clay 4.3.3.4) indicator due to late development of
foundations localised failure zone and rapid
progressive failure mechanism (refer
Section 4.3.3.4 for further discussion).
Note: *1 Most of the exceptions are discussed further in Appendix C (Section 3).
*2 All indicators includes vertical deformation at and beyond the toe (Section 4.3.3.2 ), lateral
surface displacement at the embankment toe (Section 4.3.3.3 ) and maximum lateral
displacement at depth at the embankment toe (Section 4.3.3.4).
Chapter 4: Embankments on soft ground Page 4.45

In the case of the Rio test embankment, the increase in rate of displacement at the toe
(both vertical and lateral), which occurred at about 90% of the eventual embankment
failure height, is coincident with cracking observed in the embankment. The cracking
of the embankment is considered to have had an effect on the increased rates of
displacement observed due to changes in the effective stress conditions in the
foundation. However, the lateral displacement probably caused the cracking.

4.3.4 FACTOR OF SAFETY VERSUS RELATIVE EMBANKMENT HEIGHT

The purpose of this section is to assess the relationship between embankment height and
the factor of safety for embankments on soft ground. The results (Figure 4.29) are
given in the form of factor of safety versus relative embankment height, as most of the
previously presented deformation plots are in this form.
From the database of failure case studies, the factor of safety with increasing
embankment height was taken as published for Thames (Marsland and Powell 1977),
Cubzac-A (Blondeau et al 1977) and Narbonne (Pilot 1972), and was calculated for Rio,
Muar-F, St. Alban-A and Portsmouth test embankments. The computer program
SLOPE-W was used to calculate the factor of safety of the four case studies analysed
adopting the following methodology:
• Total stress analysis using undrained strengths determined by the vane (Portsmouth,
St. Alban and Rio) or laboratory triaxial testing (Muar-F). No correction was
applied to the vane readings.
• A strength anisotropy function was used for the soft clay foundation where
sufficient data was available from the published information (Portsmouth).
• Shear strength for the crust using the median strength analysis approach proposed by
Ferkh and Fell (1994) to account for loss of strength of the crust associated with the
presence of pre-existing fissures and the cracking that occurs under the settlement of
the embankment. The crust strength is determined from assessment the profile of
liquidity index (IL) and the median of the measured vane strengths. Where the
liquidity index is greater than 1.0 the median of the vane strength is used. Above
the level where the liquidity index equals 1.0 (i.e. where IL < 1) a constant strength
equivalent to the median vane strength at the base of the layer (i.e. where IL = 1) is
adopted. The results for Portsmouth do not include this crust strength correction
because the analysis using the correction gives a factor of safety at failure of 0.9.
Chapter 4: Embankments on soft ground Page 4.46

This observation is possibly due to using an un-conservative function for the


strength anisotropy.
• Embankment strength as reported and modelling of cracking (at the location
reported) if this was observed prior to failure.
• The model was, as far as practical, allowed to determine the surface of rupture
corresponding to the minimum factor of safety. In most cases the surface of rupture
modelled was reasonably close to that observed. For the Muar-F and Rio test
embankments, difficulties were encountered in matching the observed large radius
of the surface of rupture and it was considered that this was due to the likelihood of
progressive failure at Muar-F and possibly a complex translational – rotational
surface of rupture at Rio.
• Once a factor of safety of close to one was calculated for the failure condition and a
reasonable match of the observed surface of rupture obtained, the modelling was
continued for lower embankment heights.

2.75

Narbonne
2.5
Rio
Muar-F
2.25
Thames
St. Alban-A
2
Portsmouth
Factor of Safety

Cubzac-A
1.75

1.5

1.25

0.75

0.5
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative Embankment Height, H/Hf

Figure 4.29: Factor of safety versus relative embankment height

The conclusion from the analysis is that the calculated factor of safety approaching a
failure condition is quite low, as the observed deformation behaviour at monitoring
points located within the eventual failure zone indicates. The calculated factors of
safety (Figure 4.29) fall below 1.5 at relative embankment heights of 60 to 70% of the
Chapter 4: Embankments on soft ground Page 4.47

failure height in most cases, and below 1.25 at 75 to 85% of the failure height. The
vertical and lateral displacements at and beyond the embankment toe (Figure 4.23 to
Figure 4.25) indicate, for all types of foundation conditions, an increase in the rate of
displacement with increasing embankment height typically in the range 60 to 90% of
the eventual failure height.
As shown in Figure 4.29, there is a close correlation of the plots of factor of safety
versus relative embankment height for the rotational failures despite the varying soil
types (low to high sensitivity and low to high plasticity) and embankment geometry
between the case studies. For the Thames embankment, the surface of rupture is
complex translational-rotational.

4.3.5 POST-FAILURE DEFORMATION BEHAVIOUR

4.3.5.1 Post Failure Rates of Movement


The post failure rates of movement from the case studies analysed, classified in
accordance with the velocity categories of IUGS (1995) (refer Table 1.1), are
summarised as follows:
• For St. Alban-A and James Bay (both sensitive clay foundations) the failure
occurred in minutes, with maximum rates of movement estimated at 0.9 and 1.8
m/mi n. Both classify at the upper end of the rapid category.
• For the Muar-F embankment, the failure also occurred in minutes at an estimated
maximum rate of 1 m/min, also classifying at the upper end of the rapid category. A
“fairly loud noise” (Brand and Premchitt 1989) was heard at the time of the failure
possibly indicating internal brittleness in the slide mechanism.
• For Thames and Kalix the maximum rates of movement are estimated at 30 and 45
mm/hour respectively, classifying in the moderate range. For Thames the full slip
developed over some 10 hours.
• For the Muar-EC embankment, continued movement for many days followed the
initial post failure movement of 340 mm. Ramalho-Ortigao et al (1983b) reported
the failure at Rio as “a general slip slowly occurred”. In both cases the maximum
rate of movement would probably classify within the moderate range.
Chapter 4: Embankments on soft ground Page 4.48

For the remaining case studies insufficient information was available with which to
estimate the maximum velocity of the slide mass post failure. The amo unt of
deformation and estimated velocity of the slide mass post failure for the case studies is
summarised in Table 4.1.
Where reasonable estimates of the post failure velocity are possible from the
published literature, they indicate that:
• For embankments on low plasticity, sensitive clay foundations the post failure
velocity is likely to be in the range of metres per minute (rapid category).
• For embankments on foundations of low to high plasticity clays of low sensitivity,
the post failure velocity is likely to be in the slow to moderate velocity category
(centimetres per hour, possibly up to metres/hour).

However, there are exceptions. In the case of Muar-F (a high plasticity clay
foundation of low sensitivity) the failure was at a rapid velocity, but the mechanics of
failure were thought to possibly incorporate some internal brittleness at failure (as
indicated by the large noise at failure), which contributed to the high velocity. In the
case of Kalix the post failure velocity was quite low for a sensitive clay foundation.

4.3.5.2 Analysis of the Post-Failure Deformation Behaviour


Khalili et al (1996) have proposed a simplified method for estimating the post-failure
deformation of intact sliding along a rotational surface of rupture. They applied the
method to six case studies, five were embankments on soft ground (that remained
basically intact on sliding) and one was an embankment dam subject to liquefaction
(Lower San Fernando dam).
The Khalili et al (1996) model considers two modes of failure defining the upper
and lower bounds of post-failure deformation:
• A “rapid model” based on the principle of conservation of energy where the
potential energy of the slide mass is resisted by the frictional forces along the
surface of rupture approximated using the residual shear strength. This is essentially
a simplified lumped mass model.
• A “slow model” based on equations of equilibrium between the driving force of the
slide mass and resisting forces along the surface of rupture approximated using the
Chapter 4: Embankments on soft ground Page 4.49

residual shear strength, i.e. assuming the rate of movement would be so slow that
the effect of inertia forces of the sliding mass are negligible.

The failure model is shown in Figure 4.30 and is based on the assumptions of a circular
slide surface, the failure is plane strain, the failed slide mass moves as a rigid body and
energy losses during failure are only due to the frictional forces acting along the slip
surface. Approximate solutions for the rapid model are calculated using Equation 4.6
and for the slow model using Equation 4.7, where FSresidual is the factor of safety
calculated using residual strengths along the surface of rupture, and ?i and ?f are the
initial and final positions of the centre of gravity as defined in Figure 4.30.

Rapid Model: θ f = 2θ i ( FS residual − 0.5 ) (4.6)

Slow Model: sin θ f = FS residual sinθ i (4.7)

Figure 4.30: Failure model for post-failure deformation (Khalili et al 1996)

Note that the terms “rapid” model and “slow” model defined and used by Khalili et al
(1996) are different to the definitions of “rapid” slides and “slow” slides used
throughout the thesis and the definitions of rapid and slow post failure velocity defined
by IUGS (1995).
Analysis of the post-failure deformation using the Khalili et al (1996) method has
been undertaken on four of the failure case studies of embankments on soft ground (Rio,
Chapter 4: Embankments on soft ground Page 4.50

St. Alban-A, Portsmouth and Muar-F). The results of the analysis are summarised in
Table 4.5. Details of the post failure deformation for each case study and the analysis
undertaken are given in Appendix C (Section 4).

Table 4.5: Results of the post-failure deformation analysis


Case ?i FS residual ?i – ?f ?i – ?f Calculated Observed
Study Calculated Observed Crest Crest
(º) Settlement (m) Settlement
(º) Slow Rapid (º) Slow Rapid (m)
Rio 8.5 0.43 (vane)*1 4.9 9.8 3.0 1.1 2.1 0.68
0.70 ( φ r′ )*2 2.55 5.1 0.59 1.14
Muar-F 5.1 0.40 (vane) 3.1 6.1 5.6 0.78 1.53 1.4
0.71 ( φ r′ ) 1.5 2.9 0.38 0.75
St. Alban-A 21.5 0.26 (vane) 16.1 32 8.1 2.3 4.6 1.15
3
0.75 (lab)* 5.6 10.8 0.79 1.53
Portsmouth 14.5 0.27 (vane) 10.6 21.2 10.4 3.4 6.45 3.3 – 3.4
Note: *1 vane = vane remoulded shear strength
*2 φ r′ = large strain undrained shear strength calculated from the estimated drained residual
shear strength (see below).
3
* Lab = large strain laboratory undrained shear strength

For the Rio and Muar-F test embankments, both constructed on high plasticity, low
sensitivity clay foundations, the large strain undrained shear strength was obtained by
estimating a residual friction angle ( φ r′ ) based on the clay fines content and plasticity of
the soft clay, and then calculating the large strain undrained shear strength (Sur) profile
from Equation 4.8. Estimation by this method is considered suitable only for soils that
are not contractive on shearing.

S ur = σ ′p tan φr′ (4.8)

where σ ′p is the pre-consolidation pressure of the soft clay.

The predictions based on the estimates of the residual undrained strength profile
from laboratory test results (St. Alban-A) and estimated residual friction angle (Rio and
Muar-F) are much closer to the observed post-failure deformation behaviour (Table
4.5), than the predictions using the estimates from vane testing.
Chapter 4: Embankments on soft ground Page 4.51

For the Rio test embankment, the failure occurred relatively slowly and therefore the
observed behaviour would be expected to be closer to the slow method prediction as is
the case. For St. Alban-A the failure occurred rapidly and therefore the observed
behaviour would be expected to be closer to the rapid method prediction, it is located
about mid way between the slow and the rapid model.
In the case of the Muar-F embankment, the observed settlement is significantly
greater than the prediction based on the large strain undrained shear strength obtained
by the estimate of the residual friction angle method. It is considered that the actual
surface of rupture may have incorporated a significant translational component and the
assumption therefore of rotational, intact sliding by the Khalili et al (1996) model may
invalidate its use for this embankment.
General guidelines on the use of this type of model are discussed in Chapter 5
together with the findings from analysis of failures in embankments dams and cut slopes
in high plasticity heavily over-consolidated clays (Sections 5.8.1 and 5.9.2).

4.4 P OST CONSTRUCTION BEHAVIOUR OF EMBANKMENTS ON SOFT


GROUND
The study of the post construction behaviour of embankments on soft ground (i.e. of
embankments that do not fail during construction) is limited to the following aspects:
• A discussion on the effect of the effective stress state on the post construction
deformation behaviour and excess pore water pressure response.
• The pore water pressure response and deformation behaviour in the period following
the end of construction. For a number of reported case studies it is observed that at
the end of construction excess pore water pressures continue to rise and relatively
large deformations occur. The observations from five case studies are discussed in
the context of the mechanics. This aspect of the behaviour of embankments on soft
ground is of fundamental importance, because failure may occur many days after the
end of construction as was observed at the Muar-EC test embankment which failed
some 30 days after construction.
• Estimates of long-term deformation behaviour in the form of lateral displacement at
the toe versus settlement at the centre. This is essentially a summary of the work by
Tavenas et al (1979).
Chapter 4: Embankments on soft ground Page 4.52

4.4.1 EFFECT OF EFFECTIVE STRESS STATE ON THE POST CONSTRUCTION


BEHAVIOUR

The post construction excess pore water pressure response and deformation behaviour
of a soft clay foundation at the end of construction is significantly affected by its
effective stress state in relation to the limit state surface. The two most common
effective stress states for lightly over-consolidated, soft clay foundations at the end of
construction are at a point on the yield surface below the critical state line (Figure
4.31a) and therefore in a normally consolidated condition, or else within the limit state
surface (Figure 4.31b) and therefore in an over-consolidated condition.
Assuming that the effective stress state at a point under the foundation at the end of
construction is in an over-consolidated condition (defined by point Ec in Figure 4.31b),
as drainage takes place the effective stress path will move to the right. The initial rate
of pore water pressure dissipation will be relatively rapid and the settlement relatively
small whilst the effective stress state is within the limit state surface (i.e. in an over-
consolidated condition). Once the effective stress path reaches the yield surface (Point
Y in Figure 4.31b) the foundation will be in a normally consolidated condition. The
change in the soil characteristics (increase in compressibility and decrease in coefficient
of consolidation) will result in a marked decrease in the rate of pore water pressure
dissipation and marked increase in settlement per unit increase in effective vertical
stress.
The behaviour observed under the centre of the embankment at Kars Bridge (Folkes
and Crooks 1985) presents a good example of the significant effect of effective stress
state on the post construction pore water pressure response and settlement behaviour.

4.4.2 DEFORMATION BEHAVIOUR AND PORE WATER PRESSURE RESPONSE IN THE


INITIAL PERIOD POST CONSTRUCTION

The observation of increasing pore water pressure during periods of no construction and
at the end of construction within the soft foundation of embankments constructed on
soft ground is not uncommon. Crooks et al (1984) and Becker et al (1984) reported
pore water pressure increases in 11 of 31 case studies. Of the failure case studies
analysed as part of this research, in 4 of 13 cases the pore water pressure was observed
to increase in at least one piezometer during periods of no construction or at the end of
Chapter 4: Embankments on soft ground Page 4.53

construction prior to failure. Of the remaining 9 case studies there was insufficient
information to state if pore water pressures increased during periods of no construction.

Note:
Ec = end of construction
O defines the initial effective stress state prior to
construction.
Y denotes the limit state or state boundary
surface

Figure 4.31: Effective stress paths post construction (Leroueil and Tavenas 1986).

In 6 of the failure case studies analysed, deformations (mainly lateral displacement at


the toe) were observed to increase at relatively steady rates at the end of construction
leading up to a failure condition. Significant deformations were also observed during
periods of no construction for a number of the case studies, particularly in the later
stages of construction.
The observations from five test embankments are presented (two of which failed)
and the findings from these case studies are discussed in the context of the probable
mechanics involved. For the Muar-EC, Muar-3C, Olga-C and Queensborough test
embankments details of the embankment, foundation, pore water pressure response and
deformation behaviour are given in Appendix C (Section 2). A summary of the
observations is as follows:
• The Muar-EC test embankment (constructed over soft Malaysian marine clays of
high plasticity) failed 30 days after the end of construction at an embankment height
of 4.68 m (Malaysian Highway Authority 1989a). Figure 4.32 shows that at
Piezometer P2 (located at 9 m depth below the embankment toe) the excess pore
water pressure response continued to increase after the end of construction and up to
the time of failure. The pore water pressure also increased in the non-construction
Chapter 4: Embankments on soft ground Page 4.54

period after the embankment had been constructed to 3.9 m height (97 kPa applied
load).
• At the Muar-3C test embankment (same site as Muar-EC, but the embankment did
not fail) pore water pressures in several piezometers were observed to increase for
periods of 110 to 120 days after the end of construction. During this period,
significant settlements under the centre of the embankment and lateral
displacements at the embankment toe were measured.
• The St. Alban-A test embankment (constructed over low plasticity, leached marine
clays of high sensitivity) failed some 12 to 18 hours after placement of the final fill
lift (La Rochelle et al 1974). The pore water pressure in Piezometer C1 (located at 3
m depth below the embankment slope) increased some 23 kPa over this period up to
failure (Figure 4.33).
• At the Olga-C test embankment (Lavallée et al 1992), constructed over soft varved
clays of high plasticity, pore water pressures in several piezometers were observed
to increase for periods of 50 to 70 days after the end of construction. During this
period significant vertical strains were observed in the region of the foundation
under the embankment centreline where increasing pore water pressure were
observed.
• At the Queensborough test embankment (Jardine and Hight 1987), constructed over
soft marine alluvial clays, cracking of the embankment halted construction.
Concern over potential failure of the embankment (due to rising pore water
pressures in several piezometers and continuing large deformations) led to the
reduction of the embankment height and placement of several berms at the
embankment toe.

Possible explanations of why this behaviour (rising pore water pressures and continuing
deformation at the end of construction) occurs are as follows:
• Mitchell (1986) considered it was due to structural breakdown as a result of stress
levels exceeding the pre-consolidation pressure of the soft clay, thereby leading to
greater compressibility accompanied by large decreases in permeability.
• Schiffman et al (1984) showed that large strain and self weight effects may retard
the rate of pore water pressure dissipation.
• Mesri and Choi (1979) attributed the behaviour to secondary compression effects.
Chapter 4: Embankments on soft ground Page 4.55

• Brand (1985a) referred to the unreliability of pore water pressure data usually
caused by the slow response times of piezometer systems.
• Crooks (1987) attributed the continued rise in pore water pressure to the time
dependent increase of total horizontal stress, therefore suggesting that the total stress
path changes with time. He further commented that a rising pore water pressure
condition is indicative of critical stressing of significant portions of the foundation,
but that it does not necessarily indicate an impending failure condition.

Tavenas and Leroueil (1977) considered that the limit state surface has a strain rate
or time dependency (Figure 4.34). With decreasing rates of strain in laboratory triaxial
tests they showed that the limit state surface (for Champlain Clays from the St. Alban
test site) was strain rate dependent. Other authors (Graham et al 1983; Soga and
Mitchell 1996; Vaid and Campanella 1977) have also commented on the strain rate
dependence of undrained shear strength and the limit state surface. Tavenas et al (1978)
used the YLIGHT model and concept of limit state isotaches (time dependent limit state
surfaces) to describe the creep to failure (under constant stress) behaviour of drained
and undrained triaxial samples at effective stress states above the critical state line and
below the outer limit state surface.

120

P4 (centre, 4.6 m)

100 P2 (toe, 9 m)
Excess Pore Water Pressure (kPa)

Undrained Response
Failure at
300 days
80

60

40

20

0
0 50 100 150 200 250 300 350
Time (Days)
Figure 4.32: Excess pore water pressure response with time for several piezometers at
the Muar-EC test embankment.
Chapter 4: Embankments on soft ground Page 4.56

80

Excess Pore Pressure and Fill Load (KPa) . 70


Applied Embankment Load
Piezo C1 (crest, 3.0 m) Embankment
60
Failure at 8:15
am on 13/10/72
50

40

30

20

10

0
0 1 2 3 4 5 6 7 8 9 10
Time (Days)
Figure 4.33: Excess pore water pressure response at Piezometer C1 with time at the St.
Alban-A test embankment.

In an attempt to understand the observation of increasing pore water pressure during no


construction periods, and particularly the delayed time to failure shown by a number of
the case studies, it is considered that the concept of time or strain rate dependent limit
state isotaches would predict an increase in pore water pressure with time. For
foundations in a normally consolidated condition (i.e. effective stress state coincident
with the yield surface below the critical state line) the pore water pressure would
increase with time due to contraction of the limit state surface and the requirement that
the effective stress state must be bounded by the limit state surface.
Three hypothetical situations have been considered to further explain the
observations of increased pore water pressure and delay to failure. The three
hypothetical situations are represented by effective stress states Ai, Bi and Ci (Figure
4.34), for which the outer limit state surface is representative of the end of construction
condition, the inner limit state surface representative of some time, tf, after construction
and the intermediate limit state surface representative of some intermediate time.
Assume for the moment that consolidation does not take place (i.e. the effective stress
state and limit state surface will not move to the right). The likely pore water pressure
response with time, for each hypothetical case, based on critical state soil mechanics
would be:
Chapter 4: Embankments on soft ground Page 4.57

• For stress state Ai at the end of construction, pore water pressures will increase with
time as the limit state surface contracts. This is because the effective stress state
must contract (i.e. move to the left) with the limit state surface, as states outside the
limit state do not exist. As no consolidation is assumed to occur the total stress state
will remain fixed and deformation will occur as a result of yielding. At some point
the effective stress state intersects with the critical state surface (Figure 4.34) and
with time the effective stress path follows the critical state line as is bounded by the
limit state and critical state surfaces. Therefore, strain weakening develops in the
soil described by the initial stress state Ai to the stress state Af at some time tf.
Deformations during strain weakening would occur as plastic flow and pore water
pressures would be expected to rise significantly as the total stress path is forced
downward. This behaviour (or similar type behaviour) is considered to have
occurred within the highly stressed regions of St. Alban-A and Muar-EC (amongst
other cases) where localised failure occurs and significant plastic flow develops on
the surface of rupture leading up to a failure condition.
• For stress state Bi at the end of construction, pore water pressures will increase with
time as the limit state surface contracts because, as for case Ai, the effective stress
state must contract (i.e. move to the left) with the limit state surface. As no
consolidation is assumed to occur the total stress state is assumed to remain fixed
and deformations occur as a result of yielding.
• For stress state Ci at the end of construction, the effective stress state is within the
limit state surface and therefore describes an over-consolidated condition. However,
at some intermediate time (assuming no consolidation) the effective stress state is
coincident with the limit state surface and the soil is now in a normally consolidated
condition. With further time, pore water pressures will increase as the limit state
surface contracts and the effective stress state moves to the left to position C f at
some time tf. Deformations occur as a result of yielding from the intermediate stress
state to the stress state at time tf.

It is important to note that for the hypothetical case “A” strain weakening can occur
in essentially ductile clays due to the effects of strain rate on the limit state surface.
In reality it is accepted that consolidation will occur with time from the end of
construction. Therefore, it is concluded that the processes of deformation due to
consolidation and deformation due to undrained creep (as described by the hypothetical
Chapter 4: Embankments on soft ground Page 4.58

examples) occur at the same time. Leroueil (1996) used the observations of pore water
pressure response and settlement of the Olga-C test embankment to draw similar
conclusions.

Figure 4.34: Schematic behaviour of strain rate effects on limit state surface with
decreasing strain rate in undrained creep tests (adapted from Leroueil and Marques
1996)

Comparing the case studies of Muar-EC and Muar-3C (described in Appendix C,


Section 2), the rate of increase in pore water pressure response in Piezometer 2, located
close to the surface of rupture at Muar-EC, was significantly greater at Muar-EC. In
addition, the plots of vertical inclination angle from an inclinometer at the toe of the
embankment (given in Appendix C) indicate that a localised failure condition developed
at Muar-EC (between 8 m and 9.5 m depth) whilst no localised failure condition
developed at Muar-3C. It is concluded that for Muar-EC embankment, the effective
stress path within the localised failure zone could be described by the hypothetical case
Ai to Af, and for Muar-3C at location Piezometer P2 it is more closely described by
hypothetical case Bi to Bf.
For a number of case studies analysed a significant period of time delay was
observed between end of construction and embankment failure (up to 30 days for Muar-
EC). The properties of the soil are considered to have a significant effect on the time
length of the delay. At Muar-EC the foundation is a high plasticity clay of low
sensitivity whilst at St. Alban-A and James Bay (12 to 18 hour delay in each case) the
Chapter 4: Embankments on soft ground Page 4.59

foundation is highly sensitive and brittle with significant strain weakening occurring
over relatively small strains in undrained loading. Once a localised failure condition
does occur in foundations of brittle, sensitive clays, it will progress relatively quickly
due to significant strain weakening as loads within the localised failure zone are
transferred onto the pore fluid. For clays of low sensitivity in undrained loading (i.e.
which do not contract on shearing) strain weakening and progressive failure will
develop more slowly. Therefore, due to the potentially slow development of strain
weakening and progressive failure for these foundations, the delay to failure can take
longer, as observed at the Muar-EC test embankment.
Several other issues that are important in the context of this discussion are:
• It would appear that the higher the level of shear stresses on a soil in a normally
consolidated condition the more rapid the rate and prolonged the period of pore
water pressure increase. It is considered that this is due to the greater strain rates
expected at stress states closer to the critical state line. Therefore, the limit state
surface described for higher levels of shear stress is larger than that at lower levels
of shear stress.
• Another implications of these observations at the Muar-EC and St. Alban-A test
embankments is that the factor of safety is not a minimum at the end of construction,
but is at a minimum at some time after the end of construction. Whilst for
embankments on sensitive, brittle foundations the findings indicate that this time
period is relatively short (typically a matter of hours), for clays of low sensitivity the
delay to failure (and a minimum factor of safety) may be several months.

4.4.3 LONG-TERM POST CONSTRUCTION DEFORMATION

The purpose of this section is to provide some guidelines on the lateral deformation that
occurs post construction. Methods are available for estimation of settlements under the
centre of the embankment post construction; however, estimation of the lateral
deformation is somewhat more difficult.
Tavenas et al (1979), based on analysis of some 13 case studies, proposed that the
post construction maximum lateral displacement at the embankment toe increases
linearly with the settlement under the centre of the embankment. The relationship
during the initial period of consolidation (about 5 years) is on average ∆y m = 0.16∆s

(refer Figure 4.14 for definition of terms), with a standard deviation of 0.09, and this
Chapter 4: Embankments on soft ground Page 4.60

rate may be reduced for longer periods of consolidation as observed at Skå Edeby
(Figure 4.35). They commented that the ratio ∆y m / ∆s is dependent on a number of

parameters including the factor of safety and the inclination of the embankment slope.
Figure 4.35 presents the findings from 7 case studies analysed by Tavenas et al (1979).
Figure 4.36 presents their findings for different slope geometries and stability conditions
indicating the ratio ∆y m / ∆s reduces with increasing factor of safety.

In the context of the discussion in Section 4.4.2, it would be expected that the post
construction ratio ∆y m / ∆s is also dependent on whether or not yielding and possibly

plastic flow is occurring at some depth within the foundation. The results from Muar-
EC, Muar-3C and Queensborough (Appendix C, Section 2) indicate the ratio can be as
high as 1 in the initial post construction period.

Figure 4.35: Relationship between the maximum lateral displacement at the toe, ym, and
settlement at the centre, s, at the end of construction (Tavenas et al 1979).
Chapter 4: Embankments on soft ground Page 4.61

Figure 4.36: Relationship between maximum lateral displacement (toe) and settlement
(centre) at the end of construction for different embankment geometries and stability
conditions (Tavenas et al 1979).

4.5 DISCUSSION AND CONCLUSIONS

4.5.1 INDICATORS OF AN IMPENDING FAILURE

The research on the deformation behaviour of embankments on soft ground comprised


the detailed analysis of 13 well-monitored and documented case studies from the
literature that failed during construction and an additional 4 case studies that did not fail
after construction.
The findings of the analysis of the 13 failure case studies indicates that appropriately
located deformation monitoring instrumentation can be used to identify an impending
failure condition during construction. As a general guide, for embankments constructed
on a wide variety of soft foundation conditions (clay and silts of low to high sensitivity
and low to high plasticity) the following deformation monitoring instrumentation
provides a good indication of an impending failure condition:
• Lateral surface displacement at the embankment toe. In virtually all cases analysed
the rate of lateral displacement increased significantly as a failure condition was
Chapter 4: Embankments on soft ground Page 4.62

approached, typically at embankment heights within the range 70 to 90% of the


eventual failure height.
• Vertical deformation at and beyond the toe of the embankment. The monitoring
point beyond the embankment toe must be located within the eventual failure zone.
In all cases studied it was from 2 m to 10 m beyond the embankment toe (from 0.4
to 1.5 times the eventual embankment failure height). Monitoring vertical
deformation at these locations (at and beyond the embankment toe) is not
significantly affected by settlement under the embankment load and heave,
particularly beyond the toe, can identify an impending failure condition. Heave is
generally observed at embankment heights of 60 to 90% of the eventual failure
height, with significant increases in the rate of heave movement (relative to the fill
height) at 70 to 95% of the eventual failure height.

The changes in rate and/or direction of movement described above occurs, in most
cases, beyond the threshold height of the embankment (the height at which the effective
stress state within the foundation goes from an over-consolidated to normally
consolidated condition) indicating that the observed changes in deformation behaviour
leading up to failure are not associated with changes that occur at the threshold height.
From limit equilibrium analysis it is evident that the factor of safety is typically less
than about 1.25 at embankment heights of 75 to 85% of the failure height. Therefore,
stress levels within the foundation are relatively high beyond about 75% of the eventual
failure height and are coincident with the changes in undrained deformation behaviour
observed.
Other good indicators of an impending failure condition, but which are more
specific to certain foundation conditions, are:
• Inclinometer observations at the toe of the embankment. For foundations of low
sensitivity (or ductile) clays an impending failure condition can be identified by
large plastic shear strains that occur within a localised failure zone prior to failure.
Increases in the maximum vertical inclination angle versus relative fill height are
observed at 70 to 90% of the eventual failure height.
• Cracking of the embankment is a useful guide for foundations of low sensitivity (or
ductile) clays.
• The deformation and pore water pressure response during periods of no
construction. In several case studies the pore water pressure in some piezometers
Chapter 4: Embankments on soft ground Page 4.63

was observed to increase and deformations were ongoing (i.e. creep type
movements) during periods of no construction. With increasing embankment height
(i.e. increased level of shear stress) the rate of pore water pressure increase was
more rapid and for a more prolonged period of time, and the rate and amount of
creep type deformation increased. These observations are best identified within the
foundation under the embankment slope or toe.

4.5.2 GUIDELINES ON MONITORING FOR IDENTIFICATION OF AN IMPENDING FAILURE


CONDITION

In analysing the deformation behaviour of an embankment on soft ground for


identification of an impending failure, it is considered necessary to supplement the
monitoring that identifies an impending failure condition with additional
instrumentation to better understand the mechanism of movement and therefore improve
the prediction of a failure condition. Suggested guidelines for monitoring are as
follows:
• Vertical displacement at and beyond the toe of the embankment, recommended for
identification of an impending failure;
• Lateral surface displacement at the toe of the embankment, recommended for
identification of an impending failure;
• Inclinometer installed in the embankment slope or at the toe. For all soil types the
inclinometer observations will assist in identification of the change from partially
drained over-consolidated conditions to normally consolidated undrained conditions,
as well the depth of foundation over which this change has occurred. For
foundations of low sensitivity clays the inclinometer observations will show the
development of the localised failure zone leading up to a failure condition;
• Settlement under the centre of the embankment. Whilst this is considered optional,
it is usually installed for the purposes of monitoring the amount of settlement during
construction and ongoing settlement post construction;
• Piezometers under the centreline of the embankment are considered fundamental to
the understanding of the mechanism associated with the deformation for all
foundation types. The pore water pressure response will:
− Provide an indication of the degree of consolidation whilst the foundation is in
an over-consolidated condition.
Chapter 4: Embankments on soft ground Page 4.64

− Allow estimation of the initial direction of the effective stress path and its likely
intersection point with the limit state surface, which has significant implications
for the deformation behaviour as discussed in Section 4.5.4.
− Indicate the change from partially drained over-consolidated conditions to
normally consolidated undrained conditions (i.e. when the effective stress path
intersects with the limit state surface).
• Piezometers under the embankment slope are considered optional. In several case
studies they have been located within the localised failure zone and the rate of pore
water pressure generation observed has exceeded the rate of applied loading.
However, given the generally narrow band of the localised failure zone particularly
for the sensitive foundation case studies, it is probably good fortune that a
piezometer is located within the localised failure zone.

Two additional factors that are important in the context of identification of an


impending failure condition are:
• Construction should proceed slowly in order for an impending failure condition to
be more clearly identified from the deformation monitoring. If construction
proceeds too quickly it can mask the indicators of the onset to failure. Of those case
studies for which an impending failure was clearly identified, construction rates
were less than 0.15 to 0.2 m/day. For the Muar test embankments the embankments
were constructed in stages (large lifts initially followed by lifts of lesser height) with
each stage followed by a period of no construction.
• In at least six of the failure case studies analysed, the base of the surface of rupture
was coincident (or very close to) an interface between layers in the sub-surface
profile. Possible reasons for this are considered to be due to a concentration of shear
strains at the interface of two layers with a stiffness or permeability variation, or the
presence of discontinuities at the interface.

4.5.3 POST FAILURE DEFORMATION BEHAVIOUR

It is considered that the Khalili et al (1996) model for intact sliding provides reasonable
estimates of the post failure deformation provided that the peak and residual undrained
strengths are properly estimated by laboratory triaxial tests. The “slow model” is
considered more appropriate for foundations of low sensitivity (or ductile) clays and the
Chapter 4: Embankments on soft ground Page 4.65

“rapid model” for prediction of the post-failure deformation for foundations of sensitive,
brittle cohesive soils.
The analysis undertaken was limited by the number of cases analysed and available
data for estimation of the large strain residual undrained shear strength. The analysis is
considered appropriate for rotational failures (as Khalili et al (1996) indicate). In
addition, the vane remoulded strength is considered to under estimate the actual large
strain undrained shear strength and will therefore over predict the post-failure
deformation.

4.5.4 DISCUSSION ON FAILURE MECHANISM

From analysis of the thirteen failure case studies of embankments on soft ground, it is
apparent that the case studies can be divided into three general groups:
• Sensitive, brittle clays that fail rapidly on a well-defined surface of rupture (St.
Alban, Portsmouth and James Bay).
• Sensitive clays that show a breakdown in the structure of the clay over a large zone
of the foundation prior to failure, and post failure movements are relatively slow
(Kalix and Mastemyr).
• Foundations of low sensitivity that develop significant plastic flow within a defined
zone prior to failure. In these cases progressive failure is evident at embankment
heights less than the failure height.

4.5.4.1 Sensitive Clay Foundations – Rapid Onset of Failure

The case studies St. Alban-A, Portsmouth and James Bay are considered representative
of embankments on sensitive, brittle, highly strain weakening foundations that, once a
localised failure condition is initiated, rapidly progress to a failure condition on a well-
defined surface of rupture. In all three cases the foundation comprised low plasticity,
soft to very soft, leached marine deposits.
From observations of the inclinometer at the toe of the embankment, it is apparent
that for these cases significant plastic flow does not occur until the embankment is at or
very close to its failure height. The mechanics at failure are summarised as follows:
• Once the embankment has been constructed to its failure height, creep type
undrained plastic deformations occur and excess pore water pressures increase,
Chapter 4: Embankments on soft ground Page 4.66

particularly within the zone/s of relatively high shear stress, located under the slope
of the embankment.
• In the regions of relatively high shear stress, reduction in the limit state surface due
to strain rate effects results in the effective stress state reaching the critical state line.
Once this occurs a localised failure condition is reached and with further straining
strain weakening begins to develop. These brittle materials are generally contractive
on shearing post-peak and further straining beyond the peak results in contraction of
the soil structure and transfer of load onto the pore fluid resulting in the
development of high excess pore water pressures; i.e. a static liquefaction condition.
• Due to the significant loss of strength within the localised failure zone, the size of
the failure zone expands rapidly as load is transferred onto the surrounding soil.
Thus, progressive failure develops quickly once a localised failure condition is
reached.
• The factor of safety reduces to values significantly below 1 over relatively small
shear strains giving the sliding mass significant kinetic energy at failure to reach
post failure travel velocities of metres/minute (rapid to very rapid velocity
categories).

The observations at failure of the James Bay case study (Section 4.3.3.4) confirm
the rapid development of shearing in the region of the surface of rupture leading up to
the failure and that significant shearing did not take place until shortly prior to the
failure. The observation of water rise in the inclinometer tube would be consistent with
static liquefaction on the surface of rupture due to collapse of the soil structure followed
by the rapid post failure movement of the failed embankment.
In terms of identifying an impending failure condition at embankment heights less
than the failure height, these sensitive, brittle clays are considered the most difficult to
identify due to the relatively small deformations observed and the rapid development to
a failure condition. However, it is considered that deformation monitoring at the
following locations provides a good indication of an impending failure condition:
• Vertical deformation behaviour at and beyond the toe of the embankment, in the
form of heave type movements and increases in the rate of heave movement with
increasing embankment height. Heave movements occur at St. Alban-A at about 60
to 70% of the failure height and for James Bay an increase in the rate of heave
occurs at about 75% and then again beyond 90% of the failure height (Figure 4.23
Chapter 4: Embankments on soft ground Page 4.67

and Figure 4.24). The heave movements are small, less than 15 mm up to 90 to 95%
of the failure height, but perceptible.
• Lateral surface displacement at the toe of the embankment, in the form of increasing
rates of lateral displacement with increasing embankment height. For Portsmouth
and St. Alban-A a significant increase in the rate of lateral displacement with
increasing embankment height is observed at 80 to 85% and 70 to 75% respectively
of the eventual failure height.

4.5.4.2 Sensitive Clay Foundations – “Slow” Post Failure Movement


For Kalix and Mastemyr, both highly sensitive foundations, it is considered that a
partial breakdown in the soil structure of a significant proportion of the foundation
occurred prior to reaching a failure condition. The reason for this occurrence, and
therefore the difference between this classification and the rapid onset of failure
classification for sensitive clay foundations, is considered to be related to the effective
stress path crossing the critical state line whilst still in an over-consolidated condition
(i.e. represented by path OYEc in Figure 4.8a). Given the expected significant reduction
in strength associated with the partial breakdown of the soil structure, it is not clear why
a static liquefaction and flow condition did not occur at this point. At Kalix it may be
related to the high plasticity of the sensitive clay foundation.
The occurrence of these type failures is the subject of some conjecture. Leroueil
and Tavenas (1986) consider this type of behaviour having only occurred at 2 of 45
failure case studies (including Mastemyr). Folkes and Crooks (1985) consider this type
of failure condition to be more prevalent than this.
Given that Mastemyr is a circular embankment and Kalix a rectangular
embankment, and the limited number of case studies within this group, it is difficult to
draw any firm conclusions on the prediction of the onset of an impending failure
condition. However, the indicators of an impending failure condition for case studies
where partial breakdown of the soil structure occurs are considered to include:
• Lateral surface displacement at the embankment toe;
• Vertical deformation at and beyond the toe of the embankment; and
• The development of excess pore water pressures, in piezometers under the centre of
the embankment, significantly greater than the applied vertical stress. At Kalix this
Chapter 4: Embankments on soft ground Page 4.68

was observed at embankment heights above 70 to 80% of the failure height, and at
Mastemyr above 70% of the embankment failure height.

Post failure, the rate of movement at Kalix occurred relatively slowly (45 mm/day)
in comparison to the much higher post failure velocities of St. Alban and James Bay (1
to 2 m/min).

4.5.4.3 Foundations of Low Sensitivity Clay

Eight of the thirteen failure case studies analysed (Muar-F, Muar-EC, King’s Lynn,
Narbonne, Sackville, Thames, Cubzac-A and Rio) are considered to be representative of
embankments on foundations of low sensitivity (or ductile) clays.
For foundations of low sensitivity the inclinometer observations indicate that, in
most cases, a localised failure condition initially develops at embankment heights less
than the failure height. The localised failure zone is considered to initially develop at
depth below the slope of the embankment coincident with the zone of high shear stress.
The eventual surface of rupture for these case studies was located within this localised
failure zone. It is considered that the observations of localised failure are indicative of a
progressive failure mechanism that, in general, starts at embankment heights of 70 to
90% of the eventual failure height.
The onset to failure does not occur rapidly, as for the case of sensitive, brittle soft
clay foundations (Section 4.5.4.1), for two reasons; the loss of strength from strain-
weakening is significantly less than that for the brittle, sensitive clays; and the strain-
weakening develops slowly over large strains as opposed to the rapid development over
small strains for brittle, sensitive clays. Clearly, these are typical properties that define
clays of low sensitivity.
Failure occurs when the mobilised strength on the surface of rupture is less than the
driving force. The observation of delayed failure for some of these cases is considered
to be due to undrained creep effects and the strain rate dependency of the limit state
surface. The indications of this are in the inclinometer observations and also in the
excess pore water pressure readings, which tend to increase during rest periods. The
deformation and pore water pressure observations leading up to the failure of the Muar-
EC embankment being a case in point.
Chapter 4: Embankments on soft ground Page 4.69

For the case studies of embankments on clay foundations of low sensitivity the
indicators of an impending failure condition are generally clearly identifiable. The
better indicators are considered to be:
• The maximum positive vertical inclination angle (Figure 4.27) from the
inclinometer at the embankment toe is a good indicator of an impending failure
condition. A significant increase in the vertical inclination angle was observed at 60
to 90% of the failure height for most of these case studies. Also, monitoring of the
vertical inclination angle with depth (Figure 4.26) will clearly identify the
development of the localised failure zone.
• The lateral surface displacement at the toe of the embankment (Figure 4.25) is a
good indicator for most case studies. The rate of lateral displacement increased
significantly as a failure condition is approached, typically at embankment heights
within the range 70 to 90% of the eventual failure height.
• The vertical deformation behaviour at and beyond the toe of the embankment. The
onset of heave displacements and the increase in rate of heave movements with
respect to increasing embankment height are a good indicator of an impending
failure condition.
• Cracking of the embankment was observed in four cases and is a good indicator that
the embankment is very close to failure. In the case of Muar-EC the embankment
failed some 30 days after end of construction and cracking was observed some 20
days prior to the failure.

The impending failure condition at the Rio and King’s Lynn embankments was more
difficult to identify. The above indicators were only prevalent at embankment heights
of greater than 90% of the failure height. In both cases cracks were observed in the
embankment before the deformation behaviour at the toe of the embankment indicated
an impending failure condition.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page i

TABLE OF CONTENTS

5.0 LANDSLIDES IN EMBANKMENT DAMS AND CUT


SLOPES IN HEAVILY OVER-CONSOLIDATED HIGH
PLASTICITY CLAYS ................................................................ 5.1
5.1 INTRODUCTION TO THIS C HAPTER ................................................................... 5.1
5.2 LITERATURE R EVIEW ...................................................................................... 5.2
5.2.1 Statistics on Dam Incidents Involving Slope Instability............................ 5.2
5.2.2 Progressive Failure...................................................................................5.8
5.2.3 Mechanics of Post Failure Deformation................................................. 5.11
5.3 DATABASE OF CASE S TUDIES ANALYSED ...................................................... 5.12
5.3.1 Case Studies of Failures in Dam Embankments ..................................... 5.14
5.3.2 Case Studies of Failures in Cut Slopes of High Plasticity Clays............5.15
5.4 MECHANICS OF FAILURE OF CUT S LOPES IN HEAVILY OVER-
CONSOLIDATED HIGH PLASTICITY C LAYS .....................................................5.18
5.4.1 Progressive Failure................................................................................. 5.18
5.4.2 Trigger to Failure....................................................................................5.20
5.4.3 Time to Failure........................................................................................5.22
5.4.4 Effect of Defects Within the Soil Mass .................................................... 5.23
5.5 MECHANICS OF FAILURE FOR FAILURES IN EMBANKMENT DAMS .................5.24
5.5.1 Failures in Embankment Dams During Construction.............................5.24
5.5.2 Failures in Embankments During Drawdown......................................... 5.30
5.5.3 Post Construction Failures in the Downstream Slope of
Embankments...........................................................................................5.39
5.6 POST FAILURE DEFORMATION BEHAVIOUR – F AILURES IN EMBANKMENT
DAMS ............................................................................................................ 5.44
5.6.1 Factors Affecting the Post Failure Deformation..................................... 5.44
5.6.2 Summary of Case Studies of Failures in Embankment Dams ................. 5.45
5.6.3 Failures During Construction - Post Failure Deformation
Behaviour ................................................................................................ 5.46
5.6.4 Failures During Drawdown – Post Failure Deformation Behaviour..... 5.68
5.6.5 Post Failure Deformation Behaviour of Failures in the Downstream
Shoulder After Construction.................................................................... 5.80
5.7 POST FAILURE DEFORMATION BEHAVIOUR - C UT S LOPES IN HEAVILY
OVER-CONSOLIDATED C LAYS .......................................................................5.86
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page ii

5.7.1 Summary of Factors Affecting the Post Failure Deformation ................ 5.86
5.7.2 Summary of the Post Failure Deformation and Velocity of the
Failure Case Studies ............................................................................... 5.87
5.7.3 Failures in Cut Slopes – Type 2 Slope Failure Geometry....................... 5.92
5.7.4 Failures in Cut Slopes – Type 1 Slope Failure Geometry....................... 5.94
5.7.5 Failures in Cut Slopes – Type 5 Slope Failure Geometry....................... 5.95
5.8 PREDICTION OF POST-FAILURE TRAVEL DISTANCE ....................................... 5.97
5.8.1 Khalili et al (1996) Model for Intact Slides ............................................ 5.97
5.8.2 Empirical Methods for Prediction of Travel Distance of Intact Slides.5.105
5.9 CONCLUSIONS AND GUIDELINES FOR P REDICTION OF POST-FAILURE
DEFORMATION BEHAVIOUR OF INTACT SLIDES IN SOIL SLOPES ..................5.108
5.9.1 Summary of Findings from the Case Study Analysis.............................5.108
5.9.2 Guidelines for Prediction of Post-Failure Deformation.......................5.117
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page iii

LIST OF TABLES
Table 5.1: Incidence of slope instability in embankment dams (Foster 1999) .............. 5.3
Table 5.2: Factors influencing downstream slides - accident and failure incidents
(Foster 1999) ........................................................................................................... 5.4
Table 5.3: Factors influencing upstream slides - accident and failure incidents
(Foster 1999) ........................................................................................................... 5.5
Table 5.4: Failures involving slope instability, excluding sloughing cases (Foster
1999) ....................................................................................................................... 5.9
Table 5.5: Summary of case study database of slope failures...................................... 5.13
Table 5.6: Summary of embankment dam slope stability incidents by failure slope
geometry................................................................................................................ 5.15
Table 5.7: Slope instability case studies in embankment dams by embankment
type........................................................................................................................ 5.16
Table 5.8: Mechanics of failure for failures in embankment dams during
construction. .......................................................................................................... 5.24
Table 5.9: Material types through which the surface of rupture passed of post
construction failures in the downstream shoulder................................................. 5.40
Table 5.10: Timing of the failure for post construction slides in the downstream
shoulder. ................................................................................................................ 5.42
Table 5.11: Distribution of slide volume for slides in embankment dams. ................. 5.46
Table 5.12: Summary of failure case studies in embankment dams during
construction – failures within the embankment only ............................................ 5.47
Table 5.13: Summary of failure case studies in embankment dams during
construction – failure within the embankment and foundation............................. 5.48
Table 5.14: Summary of the case studies of slides in the upstream shoulder of
embankment dams triggered by drawdown .......................................................... 5.69
Table 5.15: Post failure deformation behaviour and slope/material properties of
slides in embankment dams during drawdown. .................................................... 5.72
Table 5.16: Summary of the case studies of slides in the downstream shoulder of
embankment dams that occurred post construction. ............................................. 5.82
Table 5.17: Cut slope failures in London clay. ............................................................ 5.88
Table 5.18: Cut slope failures in Upper Lias clay........................................................ 5.89
Table 5.19: Results of the post-failure deformation analysis using the Khalili et al
(1996) method. ....................................................................................................5.102
Table 5.20: Characteristics of the case study groupings from the post failure
deformation analysis (Figure 5.52) .....................................................................5.104
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page iv

Table 5.21: Factors affecting the peak post-failure velocity for slides in
embankment dams during drawdown. ................................................................5.116
Table 5.22: Guidelines for post-failure deformation prediction using the Khalili et
al (1996) models..................................................................................................5.118
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page v

LIST OF FIGURES
Figure 5.1: Comparison of population to slope stability incidents for earthfill,
rockfill and hydraulic fill embankments (excluding sloughing and unknown
embankment zonings). ............................................................................................ 5.6
Figure 5.2: Sloughing type failure (Foster 1999)........................................................... 5.7
Figure 5.3: Sliding block model showing kinetic energy lines for different
rheological models (Golder Assoc. 1995)............................................................. 5.11
Figure 5.4: Redistribution of the potential energy after failure (D’Elia et al 1998) .... 5.12
Figure 5.5: Contours of accumulated deviatoric plastic strain, ε D P , from finite
element analysis of an excavation: 3H to 1V slope, 10 m cut slope height, Ko
= 1.5, surface suction = 10 kPa (Potts et al 1997)................................................. 5.19
Figure 5.6: Distribution of water content near the slip place from failures in cut
slopes in London clay (Skempton 1964)............................................................... 5.20
Figure 5.7: Moisture content migration to the shear zone and decreasing shear
strength with time to failure from consolidated undrained tests on brown
London clay (Skempton and La Rochelle 1965)................................................... 5.20
Figure 5.8: Annual records of cut slope failures in heavily over-consolidated clays
in the UK (James 1970)......................................................................................... 5.21
Figure 5.9: Gradual reduction in the factor of safety with time due to soil
weakening and oscillations due to climatic effects (Picarelli et al 2000) ............. 5.22
Figure 5.10: Variation of the pore water pressure coefficient, ru, with time in
cuttings in Brown London clay (Skempton 1977). ............................................... 5.23
Figure 5.11: Time to first time failure versus cut slope angle for Type 1 and Type
2 slope failure geometries in London clay. ........................................................... 5.23
Figure 5.12: Idealised development of surface of rupture during construction for
embankments with wet placed fill layer/s............................................................. 5.26
Figure 5.13: Compound type surface of rupture for failures during construction at
(a) Scout Reservation dam (Mann and Snow 1992), and (b) Waco dam
(Stroman et al 1984).............................................................................................. 5.26
Figure 5.14: Carsington dam, section through the region of the initial failure at
chainage 725 m (Skempton and Vaughan 1993). ................................................. 5.27
Figure 5.15: Carsington dam; drained strength properties of intact “yellow clay”
and solifluction shears (Skempton and Vaughan 1993)........................................ 5.28
Figure 5.16: Contours of deviatoric strain predicted by finite element analysis at
chainage 725 m prior to the failure of Carsington dam (Potts et al 1990). ........... 5.29
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page vi

Figure 5.17: Section through the failure in the upstream shoulder at San Luis dam
(courtesy of U.S. Bureau of Reclamation). ........................................................... 5.35
Figure 5.18: Drawdown causing failure and prior large drawdowns for (a) Dam
FD2, (b) Dam FD3, and (c) San Luis dam. ........................................................... 5.36
Figure 5.19: Composite shear strength envelope (Duncan et al 1990). ....................... 5.38
Figure 5.20: Effect of moisture migration on the undrained strength within the
shear band using idealised strength parameters. ................................................... 5.38
Figure 5.21: Type 1 failure slope geometries with different surface of rupture
orientations............................................................................................................ 5.45
Figure 5.22: Total post-failure deformation for slides in embankment dams during
construction. .......................................................................................................... 5.49
Figure 5.23: Maximum post failure velocity of slides in embankment dams during
construction. .......................................................................................................... 5.50
Figure 5.24: Residual drained strength from field case studies in clay soils in the
UK and ring shear tests on sand, kaolin and bentonite (Skempton 1985). ........... 5.51
Figure 5.25: Ring shear tests on normally consolidated sand-bentonite mixtures
(Skempton 1985, after Lupini et al 1981). ............................................................ 5.52
Figure 5.26: Section through Punchina cofferdam (adapted from Villegas 1982). ..... 5.54
Figure 5.27: Punchina cofferdam, (a) deformation and (b) velocity of the
downstream slope during construction (adapted from Villegas (1982))............... 5.54
Figure 5.28: Section through the failed region of Muirhead dam (Banks 1948) .........5.55
Figure 5.29: Muirhead dam, post failure deformation and velocity of the failure in
the upstream shoulder during construction. .......................................................... 5.55
Figure 5.30: Crest settlement monitoring of the slide at Waco dam (adapted from
Stroman et al 1984). .............................................................................................. 5.59
Figure 5.31: Carsington dam, deformation monitoring of upstream shoulder (a)
leading up to failure, (b) during failure (Skempton and Vaughan 1993), and
(c) the velocity of the slide mass........................................................................... 5.61
Figure 5.32: Section through failed region of Acu dam (Penman 1986) .....................5.62
Figure 5.33: Cross section of the slide during construction at Chingford dam
(Cooling and Golder 1942) ................................................................................... 5.63
Figure 5.34: Deformation and velocity of the slide at Lake Shelbyville dam
(adapted from Humphrey and Leonards (1986, 1988))......................................... 5.67
Figure 5.35: Post failure deformation and velocity of slide at Scout Reservation
dam (adapted from Mann and Snow (1992)). .......................................................5.67
Figure 5.36: Slides in embankment dams during drawdown; post failure (a)
deformation and (b) estimated velocity.................................................................5.73
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page vii

Figure 5.37: San Luis dam; the latter stages of post failure (a) deformation and (b)
velocity of the first time slide following the 1981 drawdown (data courtesy of
U.S. Bureau of Reclamation). ............................................................................... 5.76
Figure 5.38: Section through the slide at Dam FD2..................................................... 5.77
Figure 5.39: Deformation behaviour of the slide at Dam FD2 on reactivation
during the 1994 drawdown; (a) deformation and (b) velocity. ............................. 5.78
Figure 5.40: Upstream slope angle versus total post failure deformation for slides
in embankment dams during drawdown. .............................................................. 5.79
Figure 5.41: Post construction slides in the downstream shoulder; post failure (a)
deformation, and (b) estimated peak velocity. ...................................................... 5.84
Figure 5.42: Slide in the cut slope in London clay at Northolt in 1955 (Skempton
1964) ..................................................................................................................... 5.87
Figure 5.43: Post-failure slide deformation for cut slopes in London clays. ............... 5.90
Figure 5.44: Estimated peak post failure velocity of slides in cut slopes in London
clays....................................................................................................................... 5.90
Figure 5.45: Cut slope angle versus post-failure deformation at the slide toe for
cut slopes in heavily over-consolidated clays of Type 1 and Type 2 slope
failure geometry. ................................................................................................... 5.91
Figure 5.46: Cut slope angle versus estimated post-failure slide velocity for cut
slopes in heavily over-consolidated clays............................................................. 5.91
Figure 5.47: Failure in the cut slope in London clay at New Cross in 1841
(Skempton 1977)................................................................................................... 5.94
Figure 5.48: Failure in the retained cut slope in London clay at Wembley Hill
(Skempton 1977)................................................................................................... 5.95
Figure 5.49: Failure in the retained cut slope in London clay at Uxbridge in 1954
(adapted from Henkel 1956) ................................................................................. 5.97
Figure 5.50: Monitored post-failure deformation of the slide in the retained cut in
London clay at Uxbridge in 1954 (adapted from Watson 1956)...........................5.97
Figure 5.51: Failure model for post-failure deformation (Khalili et al 1996).............. 5.98
Figure 5.52: Observed deformation versus the calculated residual factor of safety. .5.103
Figure 5.53: Schematic profile of slope failure showing travel distance and travel
distance angle ......................................................................................................5.106
Figure 5.54: H/L versus the slide volume for embankme nt dams and cut slopes of
moderate to rapid post-failure velocity. ..............................................................5.107
Figure 5.55: H/L versus the tangent of the fill or cut slope angle for embankment
dams and cut slopes of moderate to rapid post-failure velocity. .........................5.108
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.1

5.0 LANDSLIDES IN EMBANKMENT DAMS AND CUT


SLOPES IN HEAVILY OVER-CONSOLIDATED HIGH
PLASTICITY CLAYS

5.1 INTRODUCTION TO THIS CHAPTER


This chapter primarily analyses the post-failure deformation behaviour of landslides in
embankment dams (excluding hydraulic fill embankment dams) and cut slopes in
heavily over-consolidated high plasticity clays. The concentration of research on slow
moving slides in these two slope types is due to the amount of published case study
data, the level of investigation undertaken on the failures and research into areas related
to landsliding in these slope types. The factors affecting the failure mechanics and post
failure deformation behaviour of slow moving landslides from this study can be
considered for the broader application to other slope types.
For the most part, landslides in embankment dams, cut slopes in high plasticity clays
and fills on soft ground remain relatively intact during failure and reach post failure
velocities ranging from very slow to rapid according to the IUGS (1995) classification
(Table 1.1 of Section 1.3), i.e. velocities from centimetres per day to several metres per
hour. These predominantly “slow” slides do not possess the mass destructive
capabilities of “rapid” slides due to their generally limited amount of post-failure
deformation and their “slow” velocity. Often, the post-failure velocity is sufficiently
“slow” that the path of the slide can be avoided or remedial measures undertaken to halt
the progress of the slide and prevent potential catastrophic consequences such as the
breach of an embankment dam. However, such slides can cause extensive property
damage on or just below the slide that can be costly to repair and result in a significant
loss of amenity, as well as the potential for litigation claims and associated indirect
loses for designers and owners.
The main objectives of the research on the post failure deformation behaviour of
embankment dams and cut slopes in heavily over-consolidated high plasticity clays are:
• Assessment of the mechanics of and trigger for failure of landslides that post failure
generally remain intact and travel at “slow” velocities;
• Assessment of the material properties and slope conditions of typically “slow” and
intact landslides;
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.2

• Assessment of the mechanics of post failure deformation behaviour affecting the


travel distance, mechanics of deformation and post failure velocity; i.e. how does
the slope failure geometry, slope angle, material properties and other factors affect
the travel of the slide mass;
• Evaluation of methods and development of new methods or guidelines (where
appropriate) for prediction of the post-failure deformation behaviour.

The chapter is structured such that the statistics on dam incidents involving slope
instability and a summary of the literature on progressive failure and the mechanics of
post failure deformation (Section 5.2) provide a background, and is followed by the
analysis of the case study database (Sections 5.3 to 5.8). The analysis is presented in
three main components: Sections 5.4 and 5.5 on the mechanics of failure; Sections 5.6
and 5.7 on the post failure deformation behaviour; and Section 5.8 on the prediction on
post failure travel distance. The prediction of post failure deformation (Section 5.8)
includes the analysis of 22 cases using the methods developed by Khalili et al (1996) for
intact sliding, and includes three failure case studies of embankments on soft ground
(from Chapter 4) and one road embankment fill slope failure. Section 5.9 presents a
summary of the research findings and guidelines for prediction of the post failure
deformation behaviour.

5.2 LITERATURE R EVIEW

5.2.1 STATISTICS ON DAM INCIDENTS INVOLVING SLOPE INSTABILITY

Foster (1999) and Foster et al (2000) present statistical data on dam failure and accident
incidents (definitions of the International Commission on Large Dams (ICOLD)). Their
database comprises some 11192 large embankment dams constructed up to 1986 (large
as defined by ICOLD). The database excludes all dams constructed in China and those
in Japan constructed prior to 1930.
Foster et al (2000) reported 124 slope instability incidents, which they categorised
into upstream and downstream incidents, and the type or location of the slide (Table
5.1). Of these 124 incidents, 12 were failures that involved a breach and uncontrolled
release of water from the reservoir. The remaining 112 were accidents.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.3

Table 5.1: Incidence of slope instability in embankment dams (Foster 1999)


Type / Location No. of Downstream Slope Upstream Slope
of Slide Incidents Incidents Incidents
Failure Accident Failure Accident
Sloughing 15 4 10 0 1
Within 60 1 33 0 26
embankment only
Through 35 1 17 0 17
embankment and
foundation
Unknown 4 5 2 1 2
Total 124 11 66 1 46

Foster (1999) and Foster et al (2000) concluded that the major factors influencing the
likelihood of downstream (Table 5.2) and upstream (Table 5.3) slope instability were:
• Slope instability is more likely in earthfill embankments than earth and rockfill
embankments. Homogeneous and hydraulic fill dam embankments are statistically
the most susceptible to slope instability.
• Slope instability is more likely for dam embankments on soil foundations than for
dam embankments on rock foundations. Of the slides passing through the
embankment and foundation, the incidence of slope instability is much greater in
foundations of high plasticity clays.
• Slope instability is much more likely in dam embankments constructed of high
plasticity clays and silts, particularly for slides in the downstream slope. Dam
embankments constructed of sands and gravels have a significantly lower incidence
of slope instability.
• Compaction, or lack there of, of the earthfill materials is also a significant factor.
The incidence of slope instability is much more likely for earthfill constructed by
placement with no formal compaction, and the least likely for rolled, well
compacted earthfill.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.4

Table 5.2: Factors influencing downstream slides - accident and failure incidents (Foster 1999)
GENERAL FACTORS INFLUENCING LIKELIHOOD OF INITIATION OF DOWNSTREAM SLIDES
FACTOR MUCH MORE MORE LIKELY NEUTRAL LESS LIKELY MUCH LESS LIKELY
LIKELY
Homogeneous Earthfill with rock toe Earthfill with filter Concrete face Zoned earthfill, zoned earth and
ZONING earthfill, earthfill earthfill, puddle rockfill, central core earth and
with corewall, core earthfill rockfill, concrete face rockfill,
hydraulic earthfill rockfill with corewall
FOUNDATION Soil foundations Rock foundations
TYPE

GEOLOGY High plasticity clays Residual soils of All other geology types
in foundation (i.e. sedimentary origin and (due to low number of
TYPES *1
marine, lacustrine) ‘soft’ sedimentary rocks foundation slide cases)
CORE Lacustrine Residual, alluvial, Glacial, aeolian
GEOLOGICAL colluvial, volcanic

ORIGIN

CORE SOIL High plasticity clays Low plasticity silts and Clayey sands (SC) Clayey gravels Silty sands and gravels (SM, GM)
and silts (for clays (ML, CL) (GC)
TYPE
rotational slides)
CORE No formal Rolled, modest control Puddle, hydraulic fill Rolled, well Rolled, well compacted (particularly
compaction (accounted for by compacted (for for embankment slides and
COMPACTION
zoning) foundation slides) sloughing)
Note: *1 For slides that incorporate the foundation
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.5

Table 5.3: Factors influencing upstream slides - accident and failure incidents (Foster 1999)
GENERAL FACTORS INFLUENCING LIKELIHOOD OF INITIATION OF UPSTREAM SLIDES
FACTOR MUCH MORE MORE LIKELY NEUTRAL LESS LIKELY MUCH LESS LIKELY
LIKELY
Homogeneous Concrete face earthfill, Earthfill with rock toe Earthfill with filter, Zoned earth and rockfill,
ZONING earthfill, Puddle core earthfill, Rockfill with corewall Zoned earthfill Central core earth and rockfill
Hydraulic fill Earthfill with corewall Concrete face rockfill
FOUNDATION Soil foundations Rock foundations
TYPE

GEOLOGY High plasticity clays Residual soils of All other geology types
in foundation (i.e. sedimentary origin and (due to low no. of
TYPES *1
marine, lacustrine) ‘soft’ sedimentary rocks foundation slide cases)
CORE Glacial, Lacustrine Alluvial Residual, Marine, Aeolian, Colluvial Glacial
GEOLOGICAL Volcanic

ORIGIN

CORE SOIL Low plasticity silts and High plasticity clays Clayey sands and Silty sands and gravels (SM,
clays (ML, CL) (CH) gravels (SC, GC) GM)
TYPE

CORE No formal compaction Rolled, modest control Puddle, hydraulic fill Rolled, well compacted Rolled, well compacted (for
(accounted for by (for foundation slides) embankment slides and
COMPACTION
zoning) sloughing)
Note: *1 For slides that incorporate the foundation
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.6

Further analysis was undertaken on slope stability incidents from the Foster database,
excluding the sloughing type failures (Figure 5.2) and cases of unknown embankment
zoning. Figure 5.1 presents a comparison of the incidence of slope instability and dam
population for various embankment types and confirms the strong weighting toward
failures in earthfill embankments, in particular homogeneous earthfill embankments,
and low incidence of instability for rockfill and zoned earthfill embankments compared
to their population. The term “rockfill” in Figure 5.1 includes zoned earth and rockfill,
and central core earth and rockfill embankment types.
The analysis also found that:
• Sliding often extends into the foundation for incidents during construction (75%,
excluding hydraulic fill dams), and for incidents in zoned earthfill embankments
(60% of incidents) and rockfill embankments (80% or 4 of 5 incidents).
• Incidents involving the embankment only (i.e. the surface of rupture is located
within the embankment only) occur mostly for failures during first filling (65% of
incidents) and in the downstream slope post construction (74% of incidents). They
also mostly occur in homogeneous embankments (70% of incidents), concrete
corewall earthfill embankments (100% of incidents) and earthfill with rock toe
embankments (70% of incidents).

40%
% of Population
35%
% of Slope Stability Incidents
(Accidents and Failures)
30%
Percentage of Total.

25%

20%

15%

10%

5%

0%
Earthfill with

Earthfill with
Rockfill
Hydraulic Fill

Concrete faced
Puddle core
Zoned earthfill

Concrete core-
Homogeneous

rockfill toe

wall earthfill
earthfill
filters

earthfill

Figure 5.1: Comparison of population to slope stability incidents for earthfill, rockfill
and hydraulic fill embankments (excluding sloughing and unknown embankment
zonings).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.7

Figure 5.2: Sloughing type failure (Foster 1999)

Excluding sloughing type failures, the overall incidence of slope instability for large
embankment dams reported by Foster (1999) is 106 incidents in 300,400 dam years up
to 1986. It is important to recognise that a small number of poorly performing
embankments have contributed significantly to the incident statistics, with 46 incidents
coming from 13 embankments.
Of the 12 slope instability related failures (i.e. incidents involving breach) reported
by Foster (1999) four were sloughing type failures (progressive sliding of the
downstream slope due to seepage through the embankment, Figure 5.2). Of the
remaining 8 failures (Table 5.4), slope instability is not directly attributable as the sole
cause of the failure incident. In two cases (Fruitgrowers Dam and Utica Dam) slope
instability was known to be the initiator of the failure incident; however, the eventual
failure was by intentional breach in the case of Fruitgrowers Dam1 and by piping in the
case of Utica Dam. In five of the eight cases slope instability was thought to be only
one of the probable causes of failure. In the case of Kaila Dam slope instability resulted
in the breach, but was preceded by excessive seepage from a leaking conduit and piping.
Therefore, slope instability on its own has not been identified as being responsible
for a failure incident. In the case of Fruitgrowers Dam it could well have occurred;
however, had the slide not damaged the outlet works it may have been possible to draw
down the reservoir in a controlled, albeit rapid, manner and averted the failure.

1
Fruitgrowers Dam (Sherard 1953) was initially constructed in 1898 and then raised in 1905, 1910 and
1935. In 1937, when the reservoir was at its highest historical level, a failure in the downstream slope
occurred which extended back to the crest of the embankment. The outlet works were inoperable as a
result of the slide and the reservoir could not be drawn down. Due to concerns over potential
uncontrolled breach as a result of back-scarp instability (the slide continued to move intermittently and
seepage was observed emanating from a high level on the downstream slope) the embankment was
intentionally breached near to the abutment.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.8

In several accidents a failure (i.e. breach) was narrowly averted; at Great Western
Dam2 the reservoir was drawn down to avoid a breach and at Lower San Fernando
Dam3 it was fortunate that the reservoir level was relatively low at the time of failure.
Foster (1999) comment that slides in dam embankments usually occur slowly thereby
providing warning of the slope stability incident and allowing time for remedial action
to avoid a breach. It is also notable that no modern dams have failed by slope
instability.

5.2.2 PROGRESSIVE FAILURE

Progressive failure is the process whereby a localised failure condition, initially


developed in part of a slope, progressively spreads through the slope culminating in
eventual slope failure. At collapse (or slope failure), the mobilised shear stress and
amount of shear strain along the surface of rupture is not uniform and overall the
average shear stress is less than the peak shear strength of the soil mass (inclusive of
defects). The degree of significance of progressive failure is related to the overall
amount of strain weakening below the peak strength of the soil mass.
Progressive failure is a significant factor associated with the mechanics of failure of
slopes in soils that show post peak strain weakening in drained or undrained loading,
including heavily over-consolidated high plasticity clays and sensitive soils (Terzaghi
and Peck 1948; Skempton 1948, 1964; Bjerrum 1967; Bishop 1967; Potts et al 1997;
D’Elia et al 1998; Picarelli 2000; Leroueil 2001).

2
At Great Western dam (Sherard et al 1963) the slide extended back into the upstream face of the
embankment. Drawing down the reservoir level and stabilising the toe region of the slow moving slide
narrowly avoided a breach. At one stage the rate of drawdown equalled the rate of vertical crest
settlement (300 mm/day) at which point the freeboard (between the crest level in the slide area and the
reservoir level) reached a minimum of 0 .9 m.
3
At Lower San Fernando dam, failure of the upstream slope occurred due to liquefaction and flow sliding
following an earthquake. At the time of the earthquake the reservoir level was relatively low. The slide
extended back into the downstream slope such that the minimum remaining freeboard was in the order of
1 m.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.9

Table 5.4: Failures involving slope instability, excluding sloughing cases (Foster 1999)
Dam Name Year Year Embankment Details Slope Slide Description of Materials Description of Failure Incident
d n
Constr Failed Type Class Height Type and Placement
(m)
Jackson’s 1930 1957 homog 0,0,0 9 down- Unk Silty sands (SM) placed by Not observed. Assumed downstream slope
Bluff stream dumping (no formal compaction) instability following wet period.
Kelly Barnes 1899 1977 other 12,x,x 6 down- Unk Earthfill placed over rock crib Attributed to failure of 1H to 1V
stream dam. downstream slope. Associated with piping
and/or localised breach.
Wheatland 1893 1969 homog 0,0,0 13 down- Unk Unknown. Originally built in Cause of failure not known. Attributed to
No. 1 (raised stream 1893, raised in 1935 and 1959. sliding of downstream slope and/or piping
1959) along conduit.
Fruitgrowers 1898 1937 homog 0,0,0 11 down- Emb & Medium plasticity clays (CL) Slide in downstream slope when reservoir at
(raised stream fndn derived from residual shales. highest level. Outlet works inoperable due
1905, 1910 Original dam placed with no to slide. Embankment intentionally
& 1935) formal compaction. breached.
Ema 1932 1940 unknown 13,x,x 18 down- Unk Unknown Slide in downstream slope. Due to piping.
stream
Snake Ravine 1898 1898 hydraulic 11,0,0 19 down- Unk Central core zone of low Slide in downstream slope, possibly
fill stream plasticity silts and clays. associated with piping.
Hydraulically placed.
Utica 1873 1902 unknown 13,x,x 21 down- Emb Unknown Slides in downstream slope. Breach
stream associated with piping following slides.
Kaila 1955 1959 concrete 9,0,x 26 upstream Unk Modest control of earthfill. Slide in upstream slope. Slide preceded by
core-wall excessive seepage around leaking conduit
earthfill and piping.
d n
Note: Year Constr = year when construction was completed Class = dam zoning classification (refer Section 1.3.3)
homog = homogeneous embankment unk = unknown Emb = failure through the embankment only
Emb & fndn = failure through the embankment and foundation
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.10

Failures in cut slopes of heavily over-consolidated high plasticity clays, where


progressive failure is significant, are prevalent in the published literature due to the
engineering significance of the structures and the delay of many years to failure after
excavation, such as cuts in London clay (Skempton 1964; James 1970) and in the
Culebra clays of the Panama Canal (Banks et al 1975). Progressive failure was also
significant for several embankment dam failures, namely Carsington dam (Skempton
and Vaughan 1993) and San Luis dam (Stark and Duncan 1991), failures in riverbank
slopes of sensitive clays (Leroueil 2001), and failures in embankments on soft ground
(Chapter 4).
The mechanics of progressive failure and the formation of a shear zone and eventual
surface of rupture are discussed in recent papers by Leroueil (2001), Picarelli (2000),
Potts et al (1997), Fell et al (2000) and Cooper et al (1998b). Several general aspects
related to progressive failure are:
• Once a localised failure condition has been reached in part of a slope, progression to
a slope failure condition does not necessarily occur. In fact, the numerical analysis
by Potts et al (1997) and research by Leroueil (2001) would indicate that in a high
proportion of slopes in brittle soils a localised failure condition has been reached in
some part of the slope.
• The progression from initiation of a localised failure condition to slope failure (or
collapse) generally requires a change in material strength properties and/or boundary
conditions (i.e. a change in stress conditions such as due erosion at the toe of the
slope). Processes leading to a reduction in material strength properties include; the
stress relief process of dilation and softening (Potts et al 1997; Bjerrum 1967), creep
under constant effective stress conditions (Tavenas and Leroueil 1977; Mitchell
1993; Leroueil 2001; Fell et al 2000), fatigue due to cyclic loading (Leroueil 2001;
Stark and Duncan 1991), weathering, and changes in pore water chemistry (Picarelli
2000).
• In cases where progressive failure is significant, the concept of the factor of safety
of a slope is not clearly defined (Terzaghi 1950; Picarelli 2000; Leroueil 2001; Potts
et al 1997), it changes with time due to strain weakening in the slope.

Further aspects of progressive failure particular to the mechanics of failure in cut


slopes in heavily over-consolidated high plasticity clays and embankment dams are
discussed in Sections 5.4 and 5.5.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.11

5.2.3 MECHANICS OF POST FAILURE DEFORMATION

Models developed for post failure deformation analysis of landslides, such as sliding
block or “lumped mass” models (e.g. Banks and Strohm 1974; Koerner 1977) and
numerical continuum models (e.g. Hungr 1995; Norem et al 1990), are based on energy
principles. The kinetic energy of the slide mass in these models is dependent on the
rheological model used to define material properties and mechanics of movement (e.g.
Figure 5.3 as an example of sliding block models), and the geometrical constraints that
define the potential energy of the slide mass at failure. Locat and Leroueil (1997)
comment that the energy balance from initial failure to travel of the slide mass is
complex and not well known, and that available models are simplistic approximations
of this behaviour.

Figure 5.3: Sliding block model showing kinetic energy lines for different rheological
models (Golder Assoc. 1995).

D’Elia et al (1998) comment that for first time slides the potential energy of the slide
mass at failure dissipates with movement into the major components of frictional energy
between the sliding mass and surface over which it slides, energy used in internal
deformation (i.e. destructuration, disaggregation and remoulding of the slide mass) and
kinetic energy (Figure 5.4). They comment, and as idealised in Figure 5.4, that the
amount of potential energy dissipated as frictional energy may vary with strain or
displacement due to strain weakening on shearing from the shear strength at the point of
failure, i.e. it is dependent on the brittleness of the material.
For soils that are not strain weakening on shearing (i.e. zero brittleness), D’Elia et al
(1998) comment that the velocity of the slide mass remains small as the potential energy
is essentially consumed in friction. For soils that are strain weakening (i.e. brittleness is
large) then the frictional energy decreases with strain and a greater part of the potential
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.12

energy is transformed into kinetic energy for movement and acceleration of the slide
mass. Picarelli (2000) also points out that the mechanics of failure are significant in the
post failure deformation behaviour.

Figure 5.4: Redistribution of the potential energy after failure (D’Elia et al 1998)

Leroueil et al (1996) and Leroueil (2001) comment that the frictional energy component
may not be easily evaluated due to the variability of the stress strain characteristics of
the material (due to variation in stress path and strength anisotropy, and variation in
effective stress conditions) and the type and development of the surface of rupture (e.g.
deformation along a well defined surface of rupture or plastic deformation within a
shear band). Leroueil (2001) comments that for structured soils the energy required for
disaggregation and remoulding of the material can have an important influence on the
post failure deformation behaviour, citing examples of failures in sensitive clays and
chalk cliffs, but that overall the literature is limited in this field.

5.3 DATABASE OF CASE S TUDIES ANALYSED


The database of failures in embankment dams and cut slopes in high plasticity clays
comprises some 91 case studies sourced mostly from published literature. For several
of the embankment dams, the data was sourced from unpublished reports obtained from
the owners or surveillance authority, and alternate names have been used for these dams
(e.g. Dam FD2).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.13

Table 5.5 presents a summary of the case study database. The failures in fills on
soft ground (analysed in Chapter 4) have been included in Table 5.5 as several of these
case studies were included in the post failure deformation analysis (Section 5.8).
For most of the case studies, the failed slide mass remained relatively intact during
sliding. Post failure deformations were generally limited to less than 2 to 5 m, but were
up to 20 to 30 m. Post failure velocities were estimated to be in the slow to moderate
classes for most cases, but several were in the rapid class.

Table 5.5: Summary of case study database of slope failures


Group / Sub- No. Volume Comments *2 Main References
Material grouping Range
Type (m3)
During 13,000 to Slope failure geometry of Types 1, 2
22
construction 1.2x106 and 5.
Embankment Mostly Type 2 slope failure Various references.
Drawdown
Dams 7000 to Refer to Appendix D
induced, 19 geometries, but several of Type 1.
(excluding 1x10 6 (Section 1) for the
upstream slope
hydraulic fill individual case study
Downstream Mostly Type 2 slope failure
dams) references.
300 to geometries. Significant number of
slope, post 13
100,000 cases in earthfill embankments
construction without embankment filters.
Type 1 * 2 7 London clays – James
Cut Slopes in (1970), Skempton
High Type 2 * 2 25 600 to
Case studies from cuts in London
(1964, 1977)
clays (25 cases) and Upper Lias
Plasticity Type 5 * 2, 40,000 * 3 Upper Lia s clays –
5 clays (12 cases).
Clays Chandler (1972, 1974),
retained cuts
James (1970)
Mainly from Canada and
Sensitive clay Scandinavian countries. Clayey silts
5 -
foundations to silty clays of high sensitivity and
high undrained brittleness. Refer to Chapter 4
Fills on Soft
(Section 4.3.1) for case
Ground * 1 Low Very soft to firm clays of low
study references
sensitivity (or undrained brittleness.
8 -
ductile) clay
foundations
Note: *1 all fill on soft ground case studies are of Type 5 failure geometry
*2 The slope failure geometries Types 1,2 and 5 are defined in Figure 1.3 in Section 1.3 of
Chapter 1.
3
* volume unknown or could not be estimated for a large number of cases
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.14

5.3.1 CASE STUDIES OF FAILURES4 IN DAM E MBANKMENTS

Some 54 case studies of slope instability incidents in dam embankments have been
analysed. Hydraulic fill embankment dams were excluded from this chapter as the
failures involved liquefaction and flow sliding, resulting in rapid to very rapid post
failure velocities (they have been included in the analysis of “rapid” landslides in
Chapter 3).
The case studies of dam embankment slope stability incidents have been classified
into three categories (Table 5.6); slides during construction, post construction slides in
the upstream slope on drawdown, and post construction slides in the downstream slope.
The weighting toward slides during construction, in comparison to their statistical
weighting in slope stability incidents in large dams (refer Section 5.2.1), is because of
the availability of slope deformation monitoring records for these slides. For the post-
construction slides, monitoring records were scarce and in most cases no records were
available.
The slope failure geometry types are given in Table 5.6. In a number of cases the
location of the surface of rupture was not precisely known.
Information on the individual case studies analysed is presented in Appendix D
(Section 1). Summary details are given in a number of tables in this chapter (refer
Table 5.6). The information presented in Appendix D includes; general details on the
dam; description of materials used in construction; description of the foundation
conditions; details of the failure (location, dimensions, post failure deformation
behaviour); comments on the hydro-geological conditions, and comments on the trigger
and mechanism of the failure.
Table 5.7 presents the classification of the slope instability case studies with respect
to embankment type. In summary:
• Earthfill embankments (homogeneous earthfill, earthfill with filters and earthfill
with rock toe) comprise 36 (67%) of the case studies analysed and make up 80% (24
of 30) of the failures through the embankment only.
• Five of the case studies were in rockfill embankments. In four of these, the surface
of rupture did not pass through the rockfill zones.

4
The term ‘failure’ here means the formation of and sliding along a distinct surface of rupture through the
soil mass. It does not mean that the dam was breached. This meaning of the term ‘failure’ is used
extensively in Sections 5.4 to 5.9.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.15

Table 5.6: Summary of embankment dam slope stability incidents by failure slope
geometry
Failure No. Location of Surface Slope Failure Location of
1 2
Category Case of Rupture * Geometry * Tabulated Data
Studies Emb Emb & Unk Type 1 Type 2 Type 5 on Individual
3 Case Studies
Only Fndn *
During 22 11 11 - 7 5 8 Table 5.12, Table
Construction (41%) poss. 9 poss. 7 5.13, Table D1.1
(Appendix D)
Drawdown 19 16 3 - 5 8 0 Table 5.14
Induced, (35%) poss. 9 poss. 13 Table D1.2
Upstream (Appendix D)
Slope
Downstream 13 3 6 4 1*4 5 2 Table 5.16
Slope, Post- (24%) poss 4 poss. 10 poss. 3 Table D1.3
construction (Appendix D)
Notes: *1 Emb = embankment only; Emb & Fndn = embankment and foundation; Unk = unknown
*2 The slope failure geometries Types 1,2 and 5 are defined in Figure 1.3 in Section 1.3 of
Chapter 1.
3
* unsure if the surface of rupture passed through the foundation
*4 failure in the downstream slope above the rockfill toe of an earthfill with rock toe
embankment dam.

5.3.2 CASE STUDIES OF FAILURES IN CUT SLOPES OF HIGH PLASTICITY CLAYS

Good quality information on the post-failure deformation behaviour of “slow” slides in


cut slopes from the literature, particularly for first time slides, is limited. What
information is available is generally anecdotal and qualitative, such as ‘x’ amount of
displacement in ‘t’ time, with very few cases having quantitative surface or sub-surface
monitoring. The approach taken for the analysis of the post-failure deformation of
“slow” slides in cut slopes has therefore been to restrict the research to soil formations
with reasonably uniform (and well published) properties, and for which the published
literature contains numerous cases of slope instability with some information on post-
failure deformation.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.16

Table 5.7: Slope instability case studies in embankment dams by embankment type
Failure Category Failure No. Homogeneous Earthfill with Earthfill Puddle core Zoned Zoned earth Concrete core -
Type *1 Cases Earthfill *2 filters *3 with rock earthfill earthfill and rockfill wall earth and
toe rockfill
Construction Emb 11 1 6 (1) 1 1 - 2*4 -
Emb & fndn 11 - 3 (1) 1 2 5 - -
Drawdown Emb 16 8 2 (1) 3 1 2 - -
4
Emb & fndn 3 1 - - - - 1* 1
Downstream Slope, Emb 3 1 1 (1) 1 - - - -
post-construction Emb & fndn 6 3 1 (1) - - 1 1*4 -
Unk 4 3 - - 1 - - -
TOTAL - 54 17 13 (5) 6 5 8 4 1
1
Notes: * Emb = failure within the embankment only; Emb & fndn = failure through the embankment and foundation; Unk = the location of the surface of rupture is not
known
*2 In 5 (of the 17) cases classified as homogeneous earthfill embankments it was unknown or not sure if filters were present.
*3 The number in brackets refers to the number of cases with foundation filters only; e.g. 6 (1) indicates six cases of failures in earthfill embankments with filters
and in one of these cases the embankment only had foundation filters.
*4 Zoned earth and rockfill dams with either thin outer rockfill zone/s (horizontal thickness less than or equal to 0.5 H, where H = the height below the dam crest)
or where the surface of rupture did not pass through the rockfill zone/s (it preferentially passed through the foundation).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.17

The cut slope failure case studies come from two geological formations, London clay
and the weathered portion of the Upper Lias clay. Both formations are located in the
UK and are derived from marine clay deposition; during the Eocene period (early
Tertiary) for the London clay, and the Jurassic period for the Upper Lias clay-shale.
Uplifting and subsequent erosion has resulted in heavy over-consolidation of both
formations. Other similarities between the London clay and weathered Upper Lias clay
are that they are both of high plasticity, typically of stiff to very stiff undrained strength
consistency and are strain weakening on shearing in both drained and undrained
loading. Further details on the material properties are given in Appendix D (Section 2).
The exposed profile of the London clay formation (Skempton 1977) typically
comprises an upper weathered zone, referred to as “brown” London clay, of 5 to 15 m
thickness overlying the unweathered “blue” London clay. The formation is heavily
fissured with no preferred orientation (Skempton et al 1969), but concentrated in the
sub-horizontal plane. Other defects (Skempton et al 1969) comprise near horizontal
bedding and jointing. The permeability of the soil mass is very low.
Chandler (1972) identifies two surface profiles within the Upper Lias clay
formation, a brecciated profile and a fissured profile. He indicated that the brecciated
profile resulted from disturbance due to permafrost during the Pleistocene age. The
depth of brecciation is about 10 m and is typified by a lack of fissuring and permeability
one to two orders of magnitude greater than for the fissured profile. The fissured profile
(Chandler 1972) of the weathered Upper Lias clay is observed where the zone of
brecciation has been eroded. Fissuring is prevalent within the upper weathered profile
and does not extend into the underlying un-weathered profile. Other defects include
near horizontal bedding. Slickensiding is observed on fissures in the more weathered
materials.
A summary of the failure case studies for cuts in London clay (25 case studies) are
given Table 5.17 and for the Upper Lias clay (12 case studies) in Table 5.18. Further
details are given in the tables in Section 2 of Appendix D. Of the failure case studies:
• Cut slope height ranged from 5 to 17 m.
• Cut slope angle ranged from as low as 13.5 degrees to as steep as 56 degrees.
• Most failures are of Type 2 slope failure geometry.
• Seven are of Type 1 failure geometry where, in most cases, the basal plane is likely
to be controlled by dominant sub-horizontal defects (bedding or jointing).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.18

• Five are of Type 5 failure geometry (all in London clay), all of which involve a
deep-seated surface of rupture passing below the lower retained portion of the cut
slope.
• Of the failures in the Upper Lias clay, eight were in the fissured profile and four in
the brecciated profile.

5.4 M ECHANICS OF F AILURE OF CUT SLOPES IN H EAVILY OVER-


CONSOLIDATED HIGH P LASTICITY CLAYS

5.4.1 PROGRESSIVE FAILURE

Progressive failure is significant in the mechanics of failure of cut slopes in heavily


over-consolidated clays. The numerical analysis of cuts in London clay by Potts et al
(1997) shows that the process of stress relief occurs concurrently with the process of
progressive failure. The process of stress relief results in time dependent dilation (or
swelling) and softening of the soil mass and changes in effective stress conditions as
excavation induced pore water pressure dissipate. The process of progressive failure is,
to a large extent, contingent on the changes in effective stress conditions and material
strength properties associated with stress relief, and results in shear induced dilation and
strain weakening of the soil mass initiating in highly shear stressed regions of the slope.
The factors affecting the effective stress conditions within the slope, and therefore
the regions of high shear stress, are the slope geometry, pore water pressure conditions
and the coefficient of earth pressure at rest, Ko (Bishop 1967; Duncan and Dunlop
1969). For a cut slope of given geometry, increasing concentration of shear stress is
observed at the toe of the cut slope with increasing Ko. Potts et al (1997) showed, by
numerical analysis using an elasto-plastic model incorporating a softening function, that
localised failure initiates in the highly shear stressed toe region of a cut slope in heavily
over-consolidated clays and then gradually spreads from the toe region into the slope
(Figure 5.5). They modelled the material properties on those of Brown London clay,
with the softening function related to the deviatoric plastic strain invariant.
The process of progressive failure involves the development of reduced pore water
pressures in the highly shear stressed region/s of the slope due to the dilative response
of the over-consolidated clay under shear. This establishes differential pore water
pressure conditions within the slope. Moisture migrates toward these regions of
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.19

reduced pore water pressure, resulting in localised dilation and softening. Evidence of
moisture migration has been reported from field investigation of cut slopes in London
clay (Figure 5.6) and laboratory undrained triaxial testing (Figure 5.7).

Figure 5.5: Contours of accumulated deviatoric plastic strain, ε D P , from finite element
analysis of an excavation: 3H to 1V slope, 10 m cut slope height, Ko = 1.5, surface
suction = 10 kPa (Potts et al 1997).

Progressive failure is significant in the mechanics of failure for cut slopes of 18 to 30


degrees in heavily over-consolidated high plasticity clays. For first time failures in cut
slopes in London clay, this is evidenced by the large time to delay to failure after
excavation, which has been reported to occur up to 50 to 100 years after excavation
(James 1970; Skempton 1977). It is also a factor, although not as significant, in the
mechanics of failure in steeper cut slopes (steeper than about 35 degrees). Failures in
steeper cut slopes in London clays are reported to have occurred within days to years
after excavation (Skempton and La Rochelle 1965; James 1970).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.20

Several other aspects of cut slope failures in heavily over-consolidated high


plasticity clays of note are the trigger for failure, time to failure and the influence of
defects. These are discussed in the following sections.

Figure 5.6: Distribution of water content near the slip place from failures in cut slopes in
London clay (Skempton 1964)

Figure 5.7: Moisture content migration to the shear zone and decreasing shear strength
with time to failure from consolidated undrained tests on brown London clay (Skempton
and La Rochelle 1965)

5.4.2 TRIGGER TO FAILURE

James (1970) commented that cut slope failures in heavily over-consolidated clays in
the UK typically occurred in the period from October to February, the cooler and wetter
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.21

months, and that the annual number of recorded slides (Figure 5.8) was greater during
years where the seasonal winter rainfall was above average. Potts et al (1997)
commented that the incidence of cut slope failures in the seasonally wetter months is
indicative of the importance of the slope hydrogeology in controlling the eventual
failure of the slope.
Chandler (1974), from observation of pore water pressure response in cut slopes in
Upper Lias clays, indicated that the seasonal fluctuation in piezometric levels was
substantial at shallow depths in the over-consolidated clay, but was very small at depths
below about 2 m. It is this fluctuation at shallow depth that is significant in the trigger
for slope failure of a large number of first time failures.
Picarelli et al (2000) comment that climate is the trigger for landsliding in slopes of
stiff clays and clay shales (Figure 5.9). They describe the factors associated with
progressive soil weakening that result in a decrease in the factor of safety of the slope
with time, as aggravating factors.

Figure 5.8: Annual records of cut slope failures in heavily over-consolidated clays in the
UK (James 1970).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.22

Figure 5.9: Gradual reduction in the factor of safety with time due to soil weakening and
oscillations due to climatic effects (Picarelli et al 2000)

5.4.3 TIME TO FAILURE

The time for equilibration of pore water pressures post excavation is considered by
Skempton (1977) to be the major reason for delay in failure of cuts in heavily over-
consolidated brown London clay. Chandler (1984a) comments that the lengthy time for
equilibration of pore water pressures is due to the very low permeability of these
materials. He adds that uniform deposition conditions over a lengthy time period,
omission of pervious horizons in the deposition sequence and high density associated
with heavy over-consolidation contribute to the low permeability.
Skempton (1977) indicates that the time to reach pore water pressure equilibrium
could take up to 50 years for cuts in London clays (Figure 5.10) and Chandler (1974)
estimates up to 60 years for cuts in the weathered, fissured Upper Lias clays. James
(1970) also reports a similar time frame for cuts in Oxford clay (heavily over-
consolidated marine clay). However, for cuts in the brecciated Upper Lias clays
Chandler (1974) estimates equilibrium is reached in less than 10 years due to the higher
permeability of the brecciated soil profile.
These findings indicate the significance of permeability of the soil mass on the time
for pore water pressure equilibration, and therefore the rate of development of
progressive failure and the time to failure.
Slope geometry is also a factor in the time to failure. Figure 5.11, for cut slopes in
London clay of Type 1 and Type 2 slope failure geometry, shows that the time to failure
after excavation increases with decreasing cut slope angle.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.23

Figure 5.10: Variation of the pore water pressure coefficient, ru, with time in cuttings in
Brown London clay (Skempton 1977).

60
Cut Slope Angle (degrees) .

50

40

30

20

10

0
0 20 40 60 80 100 120
Time to First Time Failure (years since excavation)

Figure 5.11: Time to first time failure versus cut slope angle for Type 1 and Type 2
slope failure geometries in London clay.

5.4.4 EFFECT OF DEFECTS WITHIN THE SOIL MASS

The presence of fissuring is generally synonymous with high plasticity clays formed due
to weathering processes and that are heavily over-consolidated as a result of erosion of
significant depths of overburden (typical of the marine clays in the UK). However, the
presence of defects such as fissures is not a necessary requirement for progressive
failure in cut slopes in these soil types as indicated from the numerical analysis by Potts
et al (1997).
The presence of defects though can have an effect on progressive failure depending
on the strength properties of the defect in comparison to that of the intact soil and their
orientation with respect to the orientation of the cut slope. For most defects, such as
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.24

fissuring, bedding or jointing, the modulus and strength properties are usually less than
that of the intact soil. Under the stress conditions imposed on excavation of the slope,
shear strains will therefore concentrate along these defects.
Preferential deformation along defects in the highly stressed toe region of the cut
during the early stages of progressive failure was observed at the Saxon Pit excavation
in Oxford clay (Burland et al 1977), the Selbourne cut in Gault clay (Cooper et al
1998b), and cuts in the jointed Santa Barbara clay (D’Elia et al 1998; Bertuccioli et al
1996).
D’Elia et al (1998) discuss further the effect of defects, their type and orientation, on
the shape of the surface of rupture and observed pre-failure deformation behaviour for
slopes in stiff jointed clays and structurally complex clays.

5.5 M ECHANICS OF F AILURE FOR F AILURES IN EMBANKMENT DAMS

5.5.1 FAILURES IN EMBANKMENT DAMS DURING CONSTRUCTION

Four main types of failure mechanics have been identified (Table 5.8) for failures in
embankment dams during construction.

Table 5.8: Mechanics of failure for failures in embankment dams during construction.
Failure Type No. Comments
Cases
Embankment only – undrained 5 Mainly homogeneous earthfill embankments with
failure in wet placed fill materials placed well wet of OMC (average
greater than 4% wet, but material dependent).
Embankment only – undrained 6 Undrained instability within wet, low undrained
failure controlled by wet layers in strength layers in the embankment. Mainly
fill homogeneous earthfill embankments.
Embankment and Foundation – 10 Failure controlled by weak layers in foundation.
undrained failure in low strength Either pre-sheared seams at residual strength or
foundation layer low undrained strength layers.
Embankment and Foundation – 1 Only one case, Carsington. Progressive failure
progressive failure of foundation within the over-consolidated clay foundation layer
and undrained failure in the (with solifluction shears) and undrained failure
embankment. within the embankment.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.25

5.5.1.1 Wet Placed Fills


The mechanics of failure in wet placed fills is simply undrained failure of an unstable
slope. In all five cases the earthfill within the embankment was placed well wet of the
optimum moisture content for Standard compaction, at least 4% wet. The
characteristics of the wet placed earthfill are low undrained shear strength, low pre-
consolidation pressure at compaction and development of high pore water pressures.
Even at low confining pressures these materials are near normally consolidated.

5.5.1.2 Wet Placed Layers of Fill or Low Undrained Strength Foundations


These case studies are more complex than for the failure case studies in wet placed fills.
The wet placed earthfill layers or layers of low undrained strength in the foundation
typically form the basal plane of the surface of rupture. The overall slide mechanism is
generally characterised by the development of a back-scarp and lateral shear surfaces
through partially saturated soils. Failures involving the foundations are also
characterised by the persistence of the weak foundation stratum.
The mechanism for development of a surface of rupture and kinematically
admissible slide for these failure case studies is considered to be controlled by the wet
placed earthfill layer or weak foundation layer, and to incorporate the following
process:
• Once the shear stresses imposed by the embankment exceed the undrained shear
capacity of the weak layer then plastic deformations will occur (Figure 5.12).
• With increase in the applied stresses due to embankment raising, the weak zone
cannot support the additional stresses once its stress capacity has been reached.
Therefore, the additional stresses are transferred from these highly stressed areas to
areas where the shear capacity of the soil has not yet been exceeded. This will
typically be onto the lateral margins or in the central region of the embankment
immediately above the wet placed layers (Figure 5.12), thereby concentrating
stresses in these regions.
• The third stage involves the development of a full surface of rupture as shearing
progresses into the partially saturated soils of significantly greater undrained
strength capacity and which are potentially dilative on shearing.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.26

• Once a failure surface has formed deformations are then dependent on brittleness of
the slide mechanism and materials along the surface of rupture (discussed in Section
5.6.3).

Once developed, the surface of rupture is typically a compound type rupture surface
with translational basal plane in the weak layer and a rotational back-scarp as shown in
Figure 5.13.

Figure 5.12: Idealised development of surface of rupture during construction for


embankments with wet placed fill layer/s.

Figure 5.13: Compound type surface of rupture for failures during construction at (a)
Scout Reservation dam (Mann and Snow 1992), and (b) Waco dam (Stroman et al
1984).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.27

5.5.1.3 Carsington Dam


Skempton and Vaughan (1993), amongst others, concluded that progressive failure was
a significant factor in the mechanics of failure of Carsington dam, which failed during
construction.
The central core of Carsington dam, including its unusual “boot” structure on the
upstream side (Figure 5.14), comprised high plasticity clays that were placed well wet
of Standard optimum moisture content and heavily rolled. Potts et al (1990) report the
peak undrained strength of the core as 42 kPa taking into consideration strain rate
effects, the presence of shears from trafficking and plane strain conditions, and
undrained residual strength of about 30 kPa. The pre-consolidation pressure of the core
material, estimated at about 140 to 150 kPa, is relatively low. This estimate is based on
the development of positive pore water pressures (of very high incremental pore water
pressure coefficient) at applied loads of approximately 8 m head of fill (Skempton and
Vaughan 1993) as being indicative of normally consolidated conditions.
The outer earthfill zones (Zones I and II) were of weathered mudstone placed in thin
layers and well compacted (average density ratio of 101.1% of Standard Maximum Dry
Density).
The “yellow clay” foundation, a moderately to heavily over-consolidated layer of
glacially deposited high plasticity clay, was strongly strain weakening in drained
loading beyond its peak intact strength (Figure 5.15). It contained numerous
solifluction shears formed due to glacial disturbance, the strength of which were
significantly below the peak strength of the intact clay (Figure 5.15) and along which
only small displacements were required to reach residual strength.

Figure 5.14: Carsington dam, section through the region of the initial failure at chainage
725 m (Skempton and Vaughan 1993).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.28

Figure 5.15: Carsington dam; drained strength properties of intact “yellow clay” and
solifluction shears (Skempton and Vaughan 1993)

Potts et al (1990) undertook finite element analysis of the section at which the failure
first occurred (chainage 725 m) using a non-linear elastic model with a Mohr-Coulomb
yield criterion and incorporating strain softening as a function of the deviatoric plastic
strain. The modelling showed that progressive failure was a significant factor in the
mechanics of failure. A localised failure condition was initiated in the highly stressed
lower portion of the wet placed core and boot structure when the embankment was
about 5 m below its height at failure. With further raising of the embankment the
localised failure zone propagated into the “yellow clay” foundation immediately
upstream of the core boot structure. At failure (Figure 5.16) the shear strains along the
eventual surface of rupture were not uniform. The strength conditions (Potts et al 1990)
varied from:
• In the region incorporating the lower core and boot, and section of “yellow clay”
upstream of the boot structure, the soils in the zone of high shear strain had been
sheared beyond their peak strength and strain weakened to a strength between
residual and peak.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.29

• In the head and toe regions the stress states had not yet reached failure and therefore
strengths were less than peak strength. In these regions a surface of rupture had not
yet formed when the overall embankment had reached an unstable condition.

Figure 5.16: Contours of deviatoric strain predicted by finite element analysis at


chainage 725 m prior to the failure of Carsington dam (Potts et al 1990).

Skempton and Vaughan (1993) commented that the solifluction shears within the
“yellow clay” played a significant role in the progressive failure. Under the stress
condition imposed on the foundation in the vicinity of the “boot” structure prior to
failure, the reduced strength of the solifluction shears would mean that the stress
capacity of the existing shears would be reached well before that of the intact clay. The
existing shears could not support the increasing shear stresses due to embankment
raising once their stress capacity had been reached, which would then be distributed to
the edges of the defect and supported by the intact soil. The concentration of and
increasing shear stresses at the margins of the defect as the embankment is raised will
result in localised failure of the intact soil around the defect and extension of the defect
into the intact soil. Whilst the orientation of the solifluction shears were not aligned
with the eventual surface of rupture, the overall process would have led to progressive
softening of the soil mass.
For the Carsington case study several issues are important in the context of the
failure:
• Only limited deformations were observed in the toe region of the slide prior to
failure. This is expected given that the effective stress state of the foundation in the
toe region had not yet reached the failure surface and strengths were still less than
peak strength.
• The failure was initiated in the highly stressed region of the embankment
incorporating the lower portion of the wet core, boot and “yellow clay” foundation
immediately upstream of the “boot”, as the results of the finite element analyses
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.30

indicate. This is supported by the observation of plastic deformations in the lower


portion of the core in the deepest section of the embankment from the internal
vertical settlement gauges (Skempton and Vaughan 1993).
• Restraint on the lateral margins, due the lack of persistence of the “yellow clay”
foundation and also due to changes in the topography, had a significant effect in
supporting the unstable portion of the embankment and probably in suppressing the
amount of deformation of the unstable portion of the slope prior to the actual failure
and also in the post failure stage. Once the shear capacity on the lateral margins was
reached, due to shedding of load from the failed region and also strain weakening on
the margins, the velocity of the slide mass increased and the width of the slide
increased significantly (refer Section 5.6.3.3 for discussion of the post failure
deformation behaviour at Carsington dam).

5.5.2 FAILURES IN EMBANKMENTS DURING DRAWDOWN

The failure of embankments during drawdown is complex. It is complicated by: the


embankment zoning geometry; variation in strength properties of materials under the
stress conditions imposed; permeability variation within (horizontal to vertical
permeability) and between different embankment zones; the response of the phreatic
surface under the drawdown; the drawdown itself; and progressive failure. The effects
of partial saturation and changes in degree of saturation as the wetting front progresses
within the embankment adds further complexity to some of these aspects.

5.5.2.1 Case Studies of Failures in Embankments During Drawdown


For the case studies of failure during drawdown analysed, the categories of drawdown
causing failure were:
• Failure on first drawdown, within 1 to 2 years after completion of construction (5
cases).
• Failure during the historically quickest and largest drawdown in the embankments
history (4, possibly 5 cases).
• Failure during large drawdown, but historically not the largest drawdown (5,
possibly 6 cases).
• Failure during a routine operational drawdown (4 cases).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.31

Other notable aspects of the failure case studies were:


• All but one failure occurred within 35 years of construction or raising of the
embankment, with most (16 of 19 cases) occurring within 20 years.
• Medium to high plasticity clays were the dominant material type within which the
surface of rupture was located (10 of 15 cases).
• Failures occurred in upstream slopes ranging from as steep as 1.5H to 1V
(horizontal to vertical) to flatter than 3H to 1V. The upstream slope angle was
governed to some extent by material type in the outer upstream slope.

5.5.2.2 Mechanics of Failure of Embankments During Drawdown

The mechanics of failure of embankments during drawdown is, in a number of cases,


simply a case of undrained instability of the upstream slope for the slope geometry,
material strength properties and pore water pressure conditions as a result of the
drawdown. Most of these failures are observed during the first drawdown or during the
historically largest and/or fastest drawdown. Several features of these failures are worth
highlighting:
• Failure occurred in at least two embankments (Lake Shelbyville and Old Eildon
dams) where water was impounded during the period of construction. In both cases
the slide existed prior to first drawdown, but the drawdown resulted in acceleration
of the slide mass due to removal of the water load support.
• Failures during drawdown in sandy and gravelly soils occurred during very to
extremely rapid drawdown; 2300 mm/hour over 2 hours at Forsythe dam, and > 500
mm/day at Pilarcitos dam.
• High construction pore water pressures in the outer upstream earthfill zones
(discussed below) were significant for at least three failures that occurred within 1 to
2 years of completion of construction.

A number of failures from the case studies analysed did not occur during historically
the largest and most rapid drawdown. Contributing factors to these failures include:
• Softening of over-consolidated embankment materials due to progressive strain
weakening in undrained loading (discussed below).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.32

• Softening due to saturation of partially saturated soils. An increase in saturation


with time would effectively result in a decrease of the undrained strength due to
reduction in matrix suction.
• Softening from deformation of the embankment during and post construction, e.g.
crest spreading and crest cracking on first filling and under fluctuating reservoir
operation.
• Increase in pore water pressures with time within the earthfill as the wetting front
advances under high reservoir conditions (discussed below).
• Reservoir operation in the years prior to the failure. Pore water pressures are likely
to be greater where the reservoir level has been maintained at a high level in the
years preceding a large drawdown than in the case of large seasonal drawdowns or
periods of low reservoir level in the preceding years. For most of the case studies
no information on prior operation of the reservoir is available.
• Possible breakdown of rockfill in the outer zones. This is considered to have had a
small influence in the failure case studies analysed but it may not be possible to
ignore its influence.

(a) Influence of construction pore water pressures and reservoir support


The critical period of stability for embankments constructed of wet placed earthfills or
on foundations of low undrained strength is during and shortly after construction. The
high proportion of failure case studies that occur during construction involving wet
placed earthfills or earthfill layers and weak foundations are evidence of this.
The failures in the upstream slope at Mondely, Lake Shelbyville and Dam FC10 are
considered to be cases of undrained instability due to the low undrained shear strength
of wet placed earthfill zones or wet placed earthfill layers. At Lake Shelbyville and
Dam FC10 the failure was evident during or shortly after the end of construction. In
both these cases water was impounded during the latter stages of construction and
acceleration of the slide was thought to be associated with drawdown (the acceleration
occurred within 1.5 years of the end of construction). At Mondely dam it is possible
that the drawdown triggered the slide, but it is not evident if water was impounded
during the latter stages of construction. It is possible that in all three cases the water
load on the upstream face supported the slope during the latter stages of construction,
and on drawdown this support was removed resulting in either acceleration of the
existing slide or the slide itself.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.33

The wet placed earthfills or earthfill layers in each case study were of medium to
high plasticity and were reasonably well to heavily compacted. In the case of Mondely
dam, dissipation of construction pore water pressures was very slow and they were still
evident more than 10 years after construction. This very slow dissipation of
construction pore water pressures is typical in these well compacted, medium to high
plasticity clayey soils.
The slide at Old Eildon dam also involved accelerated of the existing slide on
removal of the water load support.

(b) Time to establish equilibrium phreatic surface


The pore water pressure conditions within the embankment at the time of drawdown are
an important factor in stability under drawdown. For partially saturated soils,
advancement of the wetting front within the embankment affects the effective stress
conditions as a result of the change in pore water pressure conditions, the reduction in
undrained strength of the materials as matric suction is reduced and possibly softening
due to swelling on saturation. Therefore, the time to reach an equilibrium phreatic
surface can be a factor in the timing of instability under drawdown for embankments;
particularly for partially saturated earthfills of low permeability placed in the upstream
shoulder.
Sampna Tank, a zoned earthfill embankment with an upstream earthfill zone of
medium to high plasticity clays and upstream slope of 2H to 1V, is considered one such
example. Failures in the upstream slope occurred under drawdown 4 and 7 years after
the end of construction.
The records of pore water pressure monitoring in the upstream slope of several
embankments allow for estimation of the time to reach equilibrium conditions for
embankments constructed of low permeability fills in the upstream slope. Fell (2002)
comments that, based on his experience, a steady state condition may not be reached for
20 to 30 years after construction for large dams.
Of the failure case studies analysed, the maximum vertical depth to the surface of
rupture, where measurable, ranged from 5 to 12 m in most cases. The deep-seated
failures at San Luis dam and Old Eildon dam, both of which involved the foundation,
are exceptions. Therefore, in the majority of cases, the significant issue is the time for
development of near steady state conditions for distances of penetration (measured
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.34

perpendicular to the embankment face) up to about 10 m into the upstream face. The
results of piezometric records for three embankments analysed indicates:
• For Dam FD2 (earthfill embankment constructed of medium plasticity clays)
observation well records, although difficult to interpret, indicate relatively steady
levels in the outer earthfill (outer 10 m) within 10 to 15 years. In the central to
upstream section of the embankment significant longer time periods were required
to reach steady conditions, much greater than 20 years.
• For Dam FD3 (earthfill embankment constructed of low plasticity clays in the
failure section) pore water pressures in the outer section of the earthfill (outer 10 m)
were relatively steady some 10 years after construction. Further toward the central
embankment section pore water pressures were relatively steady some 15 to 30
years after construction.
• For San Luis dam (zoned earth and rockfill embankment with core of well
compacted, medium plasticity sandy clays) piezometer records indicate the pore
water pressures in the outer section of the broad central core (outer 10 to 20 m) were
relatively steady some 15 years after construction. Further toward the central core
section pore water pressures were relatively steady some 15 to 25 years after
construction.

Clearly three cases is not sufficient on which to base assumptions regarding pore
water pressures in the outer upstream zone of low permeability clayey earthfills.
However, all three cases do indicate that pore water pressures in the outer 10 m zone
(the zone of interest for most cases of upstream slope instability on drawdown) reach
virtually steady conditions within 10 to 15 years for clays of low to medium plasticity.
It could be tentatively concluded that the influence of rising pore water pressures
due to the advance of the wetting front have a negligible effect on the case studies
constructed of low permeability earthfills in the upstream shoulder that failed after more
than 10 years of operation.

(c) Progressive Strain Weakening in Undrained Loading


Stark and Duncan (1987, 1991) undertook back-analysis of the failure during drawdown
at San Luis dam (Figure 5.17). Their analysis incorporated a rigorous analysis of pore
water pressure response under drawdown. To achieve a calculated unstable slope
condition, drained strength properties below fully softened for the fissured foundation
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.35

(fissured due to desiccation but not slickensided) and close to (but above) fully softened
strength for the embankment materials were required. They considered that progressive
failure was a significant factor in the failure, particularly within the high plasticity
slopewash foundation, where strain weakening occurred under the higher stress
conditions imposed under large drawdown events that preceded the failure.
The drawdown resulting in failure of San Luis dam and prior large drawdowns are
shown in Figure 5.18, along with the drawdowns of Dams FD2 and FD 3. The failure at
Dam FD2 (Figure 5.18a) clearly occurred during the largest and most rapid drawdown
event in the history of the reservoir operation. At San Luis dam, the drawdown during
which failure occurred in 1981 was historically the largest and fastest, but was preceded
by a significant drawdown in 1976.
For Dam FD3, failure occurred during a relatively large drawdown in 1931 but was
preceded by drawdowns both faster in rate of drawdown and larger in depth of
drawdown. Cracking of the crest was observed in Dam FD3 following the large
drawdown in 1928, indicating that the upstream shoulder was of marginal stability
under large drawdown. No significant rainfall was recorded in the months leading up to
the slide and no rainfall was recorded in the days prior to the slide, so pressures due to
water infilled cracks did not contribute to the failure.

Figure 5.17: Section through the failure in the upstream shoulder at San Luis dam
(courtesy of U.S. Bureau of Reclamation).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.36

(a) 1685
Drawdown thought to
have caused failure

Reservoir Level (RL, m) .


1680

1675

1670

Reactivation of
1665
movement
during 1994
drawdown
1660
50 100 150 200 250 300 350
Time (Days from 1st January)
1976 1977 1988 1989
1992 1993 1994

(b) 906

904
Reservoir Level (RL, m) .

902

900

898

896

894

Failure on 2
892
August 1931
890
50 100 150 200 250 300 350
Time (Days from 1st January)
1916 1928 1930 1931

170
(c)
160 Slide during 1981
drawdown
Reservoir Level (RL, m) .

150

140

130

120

110

100
50 100 150 200 250 300 350
Time (Days from 1st January)
1975 1976 1977 1981

Figure 5.18: Drawdown causing failure and prior large drawdowns for (a) Dam FD2,
(b) Dam FD3, and (c) San Luis dam.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.37

It is considered that progressive strain weakening in undrained loading is significant in


the failures during drawdown of embankments constructed with upstream shoulders or
broad central cores of over-consolidated clay earthfills. The stress conditions imposed
under large drawdown events (preceding the failure) are considered to cause shear stress
induced reduction of pore water pressures in highly stressed regions in the embankment,
resulting in the development of differential pore water pressure conditions in the
shoulder. Migration of moisture to these regions of shear stress induced low pore water
pressure will lead to dilation and softening in undrained strength.
This progressive failure mechanism could explain why in a number of failure case
studies that failure occurred on large, but historically not the largest or most rapid
drawdown as well as the relatively low drained strengths (less than peak strength) for
earthfill obtained by back analysis of cases such as San Luis dam.

5.5.2.3 Limit Equilibrium Analysis of Slope Stability Under Rapid Drawdown

Duncan et al (1990), recognising the significance of progressive strain weakening in


undrained loading under cyclic reservoir operation, recommended the use of a
composite drained and undrained strength envelope (Figure 5.19) for assessment of the
strength properties for limit equilibrium analysis under drawdown for embankments
zones that are partly or non free draining. At low stress levels, where on shearing in
undrained loading a reduction in pore water pressures will develop, Duncan et al
recommend using the drained strength envelope, and at higher stress level, where
positive pore water pressures may develop on shearing, they recommend using the
undrained strength envelope. For the zones above the normal maximum reservoir
operating level, Duncan et al (1990) recommend using undrained strengths allowing for
the effects of softening and cracking.
As an example, consider that a compacted clay earthfill (in a homogeneous
embankment) has peak drained strength properties of c′ = 10 kPa and φ ′ = 30 degrees
and undrained strength properties (assuming saturated) approximated by Su = 75 kPa
and φu = 6 degrees (note that the undrained strength properties will vary dependent
upon the variation in Ko and Af with confining stress). The composite strength envelope
is as shown in Figure 5.20. The effective normal stress at which the envelopes intersect
is about 140 kPa. Assuming Ko = 0.7 at this effective stress, this relates to an effective
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.38

vertical stress of close to 160 kPa or roughly 16 m assuming a saturated unit weight of
20 kPa (based on effective stresses).
In this example, which is not atypical say for a well compacted rolled clay earthfill,
the zone of the upstream slope within which most slope failures occur (i.e. 5 to 12 m) is
within this over-consolidated zone of the earthfill. Thus, undrained strain weakening,
according to the concept discussed in Section 5.5.2.2, could occur under the stress
conditions associated with drawdown in the outer over-consolidated upstream slope.
The effect of this reduction in undrained strength within the developing shear band is
shown on Figure 5.20.

Figure 5.19: Composite shear strength envelope (Duncan et al 1990).

200
peak undrained strength envelope peak drained strength envelope
o o
Su = 75 kPa, phiu = 6 c' = 10 kPa, phiu = 30
Shear Stress (kPa) .

150
soils dilative on
shearing and negative
pore pressures
developed
100

50 undrained strength reduction within shear


band due to moisture content migration
and dilation under high shear stress

0
0 50 100 150 200 250 300
Normal Stress (kPa)
Figure 5.20: Effect of moisture migration on the undrained strength within the shear
band using idealised strength parameters.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.39

5.5.3 POST CONSTRUCTION FAILURES IN THE DOWNSTREAM SLOPE OF


EMBANKMENTS

Most failures in embankment dams that occur in the downstream shoulder post
construction are considered to be related to changes in effective stress conditions and
material strength properties due to changes in pore water pressure conditions, and
seepage pressures in development of the phreatic surface under the impounded reservoir
(discussed below). Consideration should also be given to the potential for progressive
failure in heavily over-consolidated earthfill and foundation materials.

5.5.3.1 Embankment Zoning, Material Type and Slope Geometry

For most of the down-slope post construction failure case studies information on the
actual slide and material properties is limited. The following discussion therefore
concentrates on the embankment types, material types and slope geometry within which
the failures occurred.

(a) Embankment type


For post construction failures in the downstream slope, the dominant embankment type
(Table 5.7) consisted of homogeneous earthfill embankments (5 possibly 7 of 13 cases).
But, the most notable observation from the failure case studies is the absence of
embankment filters or downstream more permeable zones in the embankment design
and zoning. The high proportion of failures in the downstream shoulder post
construction in these embankments types is most likely related to the high phreatic
surfaces developed within the embankment and uncontrolled seepage conditions.

(b) Location of the surface of rupture and material type


Of the thirteen failure case studies, three were known to be located within the
embankment only, six through the embankment and foundation, and in four cases it was
not known if the surface of rupture passed through the foundation. In both the zoned
earthfill and zoned earth and rockfill cases, the surface of rupture passed though the
central core and foundation.
The embankment and foundation material types through which the surface of
rupture passed were dominantly clay soils (Table 5.9). In at least eight of the thirteen
cases the embankment materials were of medium to high plasticity clays. For the slides
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.40

that were known to pass through the foundation, the foundation comprised medium to
high plasticity clays in at least four of the six cases.
The high proportion of post construction failures in the downstream shoulder that
incorporate medium to high plasticity clays is considered to be associated with the
greater loss in strength on wetting from a partially saturated condition. But, it may
simply reflect that these soils are commonly used in dam construction.

Table 5.9: Material types through which the surface of rupture passed of post
construction failures in the downstream shoulder.
Material Type Embankment Foundation
(within which the Type (where surface of rupture known
surface of rupture or possibly passed through the
passed) foundation)
High plasticity clays 4 cases (2 homogeneous earthfill, 3 cases, all failures known to be
1 zoned earthfill and 1 zoned through the foundation.
earth and rockfill)
Medium plasticity clays 4 cases (3 homogeneous earthfill 1 case, known to be through the
and one puddle core earthfill). foundation.
Low plasticity clays 1 case (homogeneous earthfill) 1 case, known to be through the
foundation.
Clays, plasticity not 2 cases (both homogeneous 4 cases, 1 known to be through
known earthfill) the foundation, 3 possibly
through the foundation.
Clayey sands 1 case (homogeneous earthfill) 1 case, possibly through the
foundation
Gravelly Soils 1 case (homogeneous earthfill) None.
(GP/GW)

(c) Downstream slope angle


The steepness of the downstream shoulder is considered to be a significant factor in the
incidence of sliding, although it is difficult to quantify due to the combined influence
and variability of material strength properties and pore water pressures. For most of the
case studies of post construction instability in the downstream slope, the down-slope
was equal to or steeper than 2H to 1V (26.6 degrees).
In the case of Arroyito dam the slope was 2.5H to 1V, however, it is considered
likely the failure was a sloughing type failure in sandy gravels with fines content of up
to 10%. In the case of the Seven Sisters dyke (zoned earth and rockfill) and Aran dam
(zoned earthfill) the failures were of Type 5 slope failure geometry through the high
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.41

plasticity clay foundation (i.e. deep seated failures where the surface of rupture
extended below and beyond the toe of the embankment). Therefore the downslope
angle was not as significant for the Type 1 and Type 2 slope failure geometries (failures
within the shoulder or possibly along the embankment / foundation interface; refer to
Figure 1.3 in Section 1.3 of Chapter 1 for definitions of slope failure geometry).

5.5.3.2 Seepage Related Failures


In nine of the thirteen case studies (Arroyito, Barton, Woodrat Knob, Fruitgrowers,
Santa Ana, Siburua, Great Western, Aran and Lake Yosemite dams), post construction
instability of the downstream slope was considered to be directly related to a high
phreatic surface and or uncontrolled seepage in the embankment or foundation.
Importantly, in at least seven of these cases the slide occurred at a time when the
reservoir level was at close to full supply level and in at least six cases seepage was
observed on the downstream slope. The likely intra and inter layer permeability
variation of the embankment materials (kh > kv, where kh = horizontal permeability and
kv = vertical permeability), and absence of embankment filters or downstream more
permeable zones to intercept lateral seepage is likely to have been a significant factor
for a number of these failure cases.
The timing of these failures (Table 5.10) was generally within 25 years of initial
construction or within 2 years of raising of the embankment. The slide at Lake
Yosemite dam (60 years after construction) is the exception in terms of timing. The
strong coincidence between the timing of the slide and high reservoir level most
probably indicates that the trigger for the slide, in most cases, is increasing pore water
pressures within the embankment and/or foundation.

5.5.3.3 Failures not Directly Attributable to Seepage

Sliding in four of the case studies (Seven Sisters, Yuba, Park Reservoir and Harrogate
dams) was considered not directly attributable to seepage and high phreatic conditions
associated with seepage of impounded water from the reservoir. The mechanics of and
trigger for failure in these cases was considered to be:
• Failure in undrained loading. At Seven Sisters dyke (Peterson et al 1957), a zoned
earth and rockfill embankment, 13 failures were recorded over a period of seven
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.42

years from the end of construction. The surface of rupture passed through the high
plasticity, wet placed core and high plasticity firm to stiff clay foundation. The
mechanism of failure is considered to be in undrained loading, and probably
influenced by the wetting up and softening of the upper fissured foundation due to
advance of the wetting front.

Table 5.10: Timing of the failure for post construction slides in the downstream
shoulder.
Timing of Slide No. Comments
(years after Cases
construction)
Within 1 year 2 Seven Sisters – surface of rupture through wet placed CH
core and soft CH foundation.
Aran dam - surface of rupture through CH core and weak
CH foundation. The slide occurred when the reservoir was
at its highest historical level.
Within 2 years (of 3 At Great Western and Fruitgrowers dams the slide occurred
recent embankment when the reservoir was at an historical high level after the
raise) recent embankment raise. Seepage on the downstream face
was observed at Yuba and Fruitgrowers.
At Yuba the failure was triggered by earthquake.
Within 5 to 10 years 3 At Woodrat Knob and Arroyito dams seepage was observed
on the downstream shoulder. At Arroyito and Siburua dams
the slide occurred (or was initiated) when the reservoir was
close to full supply level.
Within 10 to 25 2 The slide at Santa Ana and Barton dams occurred when the
years reservoir was close to full supply level. In both cases
seepage was observed on the downstream shoulder.
More than 50 years 3 The failure at Harrogate was triggered by rainfall. At
Yosemite dam seepage occurred through the foundation.
Note: CH = high plasticity clay

• Earthquake trigger. The slide at Yuba dam (homogeneous earthfill constructed of


medium plasticity sandy clays) was triggered by earthquake. It is considered that
the relatively steep downstream shoulder (slope of 1.75H to 1V) was possibly of
marginal stability prior to the earthquake, and the earthquake was sufficient to
trigger the failure. Seepage was observed to emanate from the downstream slope
prior to the raise in 1949.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.43

• Rainfall trigger. The slide at Harrogate dam (puddle core earthfill embankment) and
possibly Park Reservoir dam (homogeneous earthfill constructed of clayey sands)
were triggered by rainfall. At Harrogate dam the slide occurred during winter
following the wettest monthly (and yearly) rainfall period on record since
embankment construction. The deep shrinkage cracks in the medium plasticity
downstream shoulder fill (Davies 1953) contributed to the instability by allowing
ready access of water to the earthfill. At Park Reservoir the slide occurred in the
upper portion of the downstream slope above the rockfill toe, and seepage and
possibly rainfall may have affected the stability of the steep (1.5H to 1V)
downstream slope.

Degree of compaction of earthfill materials is also an important factor for


consideration in the stability of the downstream slope. Saturation of poorly compacted
earthfills initially placed in a partially saturated condition will result in a large loss in
undrained strength. The deformation behaviour at the Hume Dam No. 1 embankment
(Cooper et al 1997) is an example of the effect of saturation of poorly compacted
earthfill. The earthfill shoulders of the concrete corewall embankment, constructed
from 1921 to 1936, were of low to medium plasticity sandy clays and clays, and were
poorly compacted by horses hooves and cartwheels. Investigations of the downstream
slope in the 1990’s, undertaken due to concern over the large observed deformations,
indicated that saturation of the lower portion of the downstream earthfill had reduced
the undrained strength consistency from an initially very stiff condition when partially
saturated to much lower strengths consistent with normally consolidated behaviour.
Cooper et al (1997) considered that the saturation occurred due to seepage and
leakage following construction. Their analysis showed that the embankment
deformations observed and low factor of safety of the downstream slope at periods of
high reservoir level were due to saturation and the resultant strength reduction in the
lower portion of the downstream earthfill. Undrained strain weakening on shearing in
the saturated earthfill layer was used in their numerical model to explain the observed
deformation behaviour. When residual strength was used within this saturated layer in
the downstream shoulder, their limit equilibrium analysis showed the factor of safety of
the approached unity under full supply conditions.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.44

5.6 P OST F AILURE D EFORMATION BEHAVIOUR – F AILURES IN


EMBANKMENT DAMS
This section presents the post failure deformation behaviour for the case studies of
failures in embankment dams. Data on travel distance and travel velocity of the case
studies is presented, and the factors affecting the deformation behaviour of the slide
mass are discussed. Greater detail is given for several case studies to highlight some of
the factors affecting the post failure deformation behaviour.

5.6.1 FACTORS AFFECTING THE POST FAILURE DEFORMATION

As a prelude to presentation of the case study data, several important factors affecting
the post failure deformation of the case studies (both embankment dams and cuts in
heavily over-consolidated high plasticity clays) are summarily introduced. Further
discussion on their influence is given in the following sections and in Section 5.7 for
cuts in heavily over-consolidated high plasticity clays. The factors are:
(i) The potential for strain weakening of the soil on shearing post failure, as
discussed in Section 5.2.3.
(ii) Slope failure geometry. The slope failure geometry directly affects the potential
energy of the slide mass. Type 1 geometries have greater potential energy than
Type 2 geometries and in turn Type 5 geometries (refer Figure 1.3 in Section 1.3
of Chapter 1 for slope failure geometry definitions).
(iii) Orientation of the surface of rupture. For a given material and given slope failure
geometry the orientation of the surface of rupture has a significant effect on the
post failure deformation behaviour. For example, a Type 1 geometry with steep
basal angle of the surface of rupture (Figure 5.21a) has the potential for greater
travel distance at higher velocity than for a surface of rupture with near horizontal
basal angle (Figure 5.21b). In the former case the slide has the potential to realise
its potential energy, and in doing so a significant amount of the potential energy
may be dissipated as kinetic energy resulting in acceleration of the slide mass and
relatively large travel distance. In the latter case the slide may not realise its
potential energy and most, if not all, of the slide mass may remain in the source
area and only travel a short distance.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.45

(iv) Brittleness associated with the slide mechanism (e.g. internal brittleness or toe
buttressing) and release of the slide on the lateral margins. Some of the potential
energy that becomes available from these sources will be redistributed to kinetic
energy resulting in acceleration of the slide mass.

Figure 5.21: Type 1 failure slope geometries with different surface of rupture
orientations.

5.6.2 SUMMARY OF CASE STUDIES OF FAILURES IN EMBANKMENT DAMS

Of the fifty-four embankment dam case studies, the failures during construction
generally have better quality and more detailed information on material properties and
deformation behaviour. This is due mainly to the early detection of the failure from
observation of warning signs (e.g. cracking) or monitoring, and then further close
observation and monitoring as the slide progressed, as well as the level of investigation
generally undertaken post failure. Therefore, discussion of the post failure deformation
behaviour and on the factors influencing the deformation is more substantial for the
failures during construction than for the failures during drawdown and post construction
in the downstream shoulder.
Table 5.11 presents a distribution of slide volume for the case studies. The failures
during construction are generally of greater volume and the post construction slides in
the downstream slope generally of the smallest volume. This is reflected in the
typically greater slide depth and width of the failures during construction.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.46

Table 5.11: Distribution of slide volume for slides in embankment dams.


Slide Volume (where known or possible to estimate)
No of
Timing of Slide = 5,000 m3 > 5,000 and > 50,000 and > 500,000 m3
Cases
= 50,000 m3 = 500,000 m3
During 22 none 5 (of 15) 5 (of 15) 5 (of 15)
construction 0% 33% 33% 33%
Post-construction, 19 none 6 (of 9) 2 (of 9) 1 (of 9)
on drawdown 0% 67% 22% 11%
Post-construction, 13 9 (of 12) 1 (of 12) 2 (of 12) none
downstream slope 75% 8% 17% 0%

5.6.3 FAILURES DURING CONSTRUCTION - POST FAILURE DEFORMATION BEHAVIOUR

Table 5.12 and Table 5.13 present a summary of the embankment geometry, material
properties and deformation behaviour of the twenty-two case studies of slope instability
in embankment dams during construction; Table 5.12 for failures within the
embankment only and Table 5.13 for failures through the embankment and foundation.
Of the failure case studies, nineteen were observed during the placement of fill materials
and the other three cases were not observed or recognised until shortly after
construction had been completed (Lake Shelbyville, Dam FC10 and Scout Reservation
dams). In addition, three of the case studies (Otter Brook, Skiatook and Truscott dams)
were not considered as failures and have been classified as cases of lateral bulging due
to the low moduli of wet placed earthfills.
The post failure deformation and estimated maximum velocity of the slide mass for
the failures during construction are presented in Figure 5.22 and Figure 5.23
respectively. Where possible the velocities have been determined from reported
deformation monitoring results, but in a number of cases has been estimated based on
the amount of deformation and qualitative information on the times over which the
deformation occurred. In general, these latter estimates are broad range as indicated by
the range given in Figure 5.23.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.47

Table 5.12: Summary of failure case studies in embankment dams during construction – failures within the embankment only
Embankment Geometry Material Properties Summary of Post Failure
Slope Const n
Name of Year of (at failure) (in zone of the surface of rupture) Deformation
Failure Rate, Comments
Dam Failure Emb Type Hgt Slope ASCS * 1 Comp n Moisture Deformn Velocity *4
Geometry (mm/day)
(m) (o) *5 Class n Rat g *2 (%) * 3 (m) (mm/day) Category
Punchina 1980 Type 1 E’fill – rock 45 26.6 570 to 1500 SM – silty sand 2 4.1% wet 1.57 250 to 350 slow Lateral bulging of downstream shoulder
toe (2,0,0) (d) during construction.
Troneras 1962 Type 1 E’fill - filters 31.5 18.4 360 to 1330 ML –sandy silt, med. 2 4.0% wet unk 600 Slow to Lateral bulging of downstream shoulder
(1,1,1) (d) plasticity (high ?) moderate during construction.
Muirhead 1941 Type 1 Puddle E’fill 21.3 21.8 85 at failure CL – med. plasticity 1 well wet of 2.1 40 to 115 Slow Failure of upstream shoulder
(8,0,1) (u) sandy clay core and OMC incorporating core and upstream
CL/SC shoulders. earthfill.
Waghad 1884 Type 2 (?) Homog (0,0,0) 32 26.6 240 CH/MH – high 1 6% wet 15.2 unk Moderate Failure in upstream shoulder during
(u) plasticity reconstruction of gorge section.
Dam FC10 1937 Type 1 Zoned E&R 19.5 22 avg 60 CH earthfill, SC/SM 2 4.5% wet 2.8 unk Slow to Failure in upstream shoulder shortly
(4,0,0) (u) random fill and outer moderate after end of construction. Affected by
rockfill. reservoir level fluctuation.
Otter Brook 1957 Type 1 E’fill - filters 40.5 21.8 470 (max.) SC – clayey fine sand 3 0.6% dry 0.88 40 slow Lateral bulging of upstream shoulder
(1,1,1) (u) (20% > 1% during construction.
wet)
Truscott 1981 Type 1 E’fill - filters 30 18.4 360 (max.) C? – clayey, unk Spec ± 2% 0.84 20 Slow Lateral bulging of downstream shoulder
(1,1,1) (d) plasticity unk. (3 ?) during construction.
Skiatook 1982 Type 1 E’fill - filters 33.8 16 230 C? – clayey, unk Spec ± 2% 0.79 23 Slow Lateral bulging of downstream shoulder
(1,1,1) (d) plasticity unk. (3 ?) during construction.
Lake 1970 Type 1 E’fill - filters 33 21.8 100 CL – sandy clay, 3 0.1% wet 0.39 24 Slow to Failure in upstream shoulder after the
Shelbyville (1,1,1) (u) med. plasticity very slow end of construction. Affected by
reservoir level fluctuation.
Scout 1984 Type 2 E’fill - filters 13.7 26.6 30 (?) CL – silty clay, low unk unk 1.8 10 Slow Failure in upstream shoulder after the
Reservation (1,0,1) (u) to med. plasticity end of construction. Latter part of
(3 ?) deformation affected by infiltration of
rainfall into cracks.
Acu 1981 Type 2 Zoned E&R 34.8 21.8 unk CH (?) earthfill in unk Lower wet, 25 1.2 x 106 Rapid Failure in upstream zoned earthfill
(4,1,1) (u) core and under then ± OMC shoulder. Passed through wet clay fill
upstream rockfill. (3 ?) zone under upstream shoulder.
Note: Notes to Table 5.12 follow Table 5.13.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.48

Table 5.13: Summary of failure case studies in embankment dams during construction – failure within the embankment and foundation
Embankment Geometry Earthfill Properties (in Summary of Post Failure
Slope Foundation Properties
Name of Year of (at failure) zone of surface of rupture) Deformation
Failure 1, 3 n (through which surface Comments
Dam Failure Emb Type Hgt Slope Description * * Comp Deform
n Velocity *4
Geometry of rupture passed)
(m) (o) *5 Rat g *
2
(m) (mm/day) Category
Waco 1961 Type 5 E’fill - filters 25.9 19.7 CL (?) – sandy and 3 CH - bentonitic clay bedding 1.57 250 to slow Deep seated compound slide in the
(1,1,1) (d) shaly clays. defect, pre-sheared 350 downstream shoulder.
Dam FC21 1969 Type 5 Zoned E’fill 43 18.4 SC/GC – clayey sands 3 CH - bentonitic shear zones, 0.9 unk Slow to Lateral spreading of foundation, no
(3,1,1) and gravels, 0.2% dry pre-sheared. very slow surface expression of deformation.
OMC
West Branch 1964 Type 5 E’fill - filters 21.5 14 CL (?) – “impervious unk CL – silty clays of low to 0.76 21 Slow Deep seated failure in upstream
(1,0,1) (u) fill” (3 ?) medium plasticity, soft to stiff. shoulder, limited to closure section.
North Ridge 1953 Type 5 E’fill - filters 18.3 12.5 CL – low plasticity 3 CH – high plasticity clays, firm 1.37 unk Moderate Deep seated compound failure in
(1,1,1) (d) clays, 0.1% dry OMC to stiff. (?) downstream shoulder.
Marshall 1930 Type 5 24.4 20.5 unk CL/CH (?) – silty clays, variable 15 6 Rapid
E’fill – rock CL (?) – “impervious 1.5 x 10 Deep seated compound failure in
Creek toe (2,0,0) (d) fill” (2 or 3 ?) consistency, soft to firm (?) downstream shoulder. Possible flow
liquefaction of loessic soils in
foundation.
Chingford 1937 Type 2 Puddle E’fill 7.9 21.8 CH clays for core and Puddle CH – high plasticity, soft to 4.3 unk moderate Compound failure in downstream
(8,0,1) (d) shoulder fill. (1 outer) very soft. shoulder.
Hollowell 1937 Type 5 Puddle E’fill 11.4 14 CH for core, sandy soils Puddle CL/CH (?) – soft to very soft 0.49 unk Slow Deep seated compound failure in
(8,1,1) (d) in shoulder. (1? outer) clay layer in foundation. downstream shoulder.
Lafayette 1928 Type 5 Zoned E’fill 35.5 20 CL/CH (?) – clayey 2 to 3 CH – high plasticity sandy clays 10 2000 Moderate Deep seated compound failure in
(3,0,0) (d) soils placed dry of downstream shoulder.
OMC
Galisteo 1970 Type 5 Zoned E’fill 38.5 17.5 CL – low to medium 3 CL/CH (?) – clay stratum, firm 2 50 Slow Deep seated compound failure in
(3,1,1) (u) plasticity sandy clays, to stiff. upstream shoulder.
placed ± OMC
Park 1981 Type 2 (?) Zoned E’fill 21.3 19.7 CL (?) – lacustrine unk CL (?) – lacustrine varved clays 0.9 unk Slow to Compound failure in upstream shoulder.
Reservoir (3,1,1) (u) clays, shoulder is of and silts. moderate
sandy moraine
Carsington 1984 Type 2 Zoned E’fill 28 18.4 CH core zone (8% wet 3 CH – high plasticity clays, 18.1 17000 to Moderate Compound failure in upstream shoulder.
(3,0,1) (u) of OMC) with heavily over-consolidated, 24000 Passed through wet central core and clay
shoulders of weathered solifluction shears. foundation.
mudstone (SC/CL)
Note: See next page for notes to Table 5.13.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.49

Notes for Table 5.12 and Table 5.13:


Emb = embankment Hgt = height Class n = classification
Const n = construction Comp n Ratg = compaction rating (see *2)
Homog = homogeneous earthfill embankment
E’fill – filters = earthfill embankment with filters
E’fill – rock toe = earthfill embankment with rock toe
Puddle E’fill = puddle core earthfill embankment
Zoned E’fill = zoned earthfill embankment
Zoned E&R = zoned earth and rockfill embankment
OMC = Standard optimum moisture content
*1 Soil classification to the Australian soil classification system (AS 1726-1993)
*2 Compaction Rating, 1 = no formal compaction, 2 = rolled moderate compaction, 3 = rolled
well compacted.
3
* = moisture contents of earthfill are relative to Standard optimum moisture content (OMC).
*4 = IUGS (1995) velocity class, refer Table 1.1 in Section 1.3 of Chapter 1.
*5 (d) = downstream slope, and (u) = upstream slope.

100.0
high plasticity clays and silts low to medium or unknown sandy
emb emb and fndn plasticity clays and silts soils
Total Deformation (m)

10.0

1.0

0.1
Dam FC10

Dam FC21

Muirhead
Acu Dam

Hollowell

Truscott Dam

Skiatook Dam

Galisteo Dam
Lafayette Dam

Otter Brook
Waco Dam

Troneras Dam
North Ridge

Carsington
Chingford

Lake Shelbyville

Scout Reservation

Punchina
Park Reservoir
Waghad

Marshall Creek
West Branch

Figure 5.22: Total post-failure deformation for slides in embankment dams during
construction.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.50

1.E+07 low to medium or sandy


high plasticity clays and silts unknown plasticity clays soils
1.E+06
Rapid
Maximum Velocity (mm/day) .

1.E+05

1.E+04
?
? Moderate
?
1.E+03

? ? ?
? ?
1.E+02
Slow

1.E+01

Very slow
1.E+00
Dam FC10

Dam FC21

Galisteo Dam
Lafayette Dam

Scout Reservation
Hollowell

Truscott Dam

Skiatook Dam
Acu Dam

Punchina
Waco Dam

North Ridge

Carsington

Lake Shelbyville

Park Reservoir
West Branch
Muirhead

Otter Brook
Marshall Creek
Chingford

Troneras Dam
Waghad

Figure 5.23: Maximum post failure velocity of slides in embankment dams during
construction.

5.6.3.1 Slope Failure Geometry and Orientation of the Surface of Rupture


Slope failure geometry was found not to be a significant factor in the post failure
deformation behaviour for the failures in embankment dams during construction. This
was considered to be due to the typical compound orientation of the surface of rupture
of most slides during construction for the Types 1, 2 and 5 slope failure geometries,
influenced by either weak layers in the foundation or wet, low undrained strength layers
within the earthfill. These weak or wet layers formed a near horizontal translational
basal component of the surface of rupture typified by Scout Reservation dam (Figure
5.13a), Waco dam (Figure 5.13b) and Carsington dam (Figure 5.14).
For some of the failures in wet placed earthfills, such as Punchina, Troneras,
Waghad and Dam FC10, the surface of rupture was considered likely to be closer to
rotational than compound. The failures at Punchina and Troneras, within wet placed
silty sands and sandy silts, were relatively shallow. In contrast, the failures in high
plasticity clays (Waghad and Dam FC10) were relatively deep-seated rotational slides.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.51

5.6.3.2 Material Type


Material type can have a significant affect on the post failure deformation behaviour.
Figure 5.22 and Figure 5.23 show that, with exceptions, greater deformations and higher
velocities were generally observed for failures where the surface of rupture passed
through high plasticity clays. For the low to medium plasticity clays and silts, and
sandy soil types deformations were of lesser magnitude, generally limited to less than 2
to 2.5 m, and post failure velocities were generally in the slow class.
For most of the case studies of failure during construction the mechanics of failure
(Section 5.5.1) involved shearing through low undrained strength, and likely near
normally consolidated layers, in either the foundation or earthfill. For these near
normally consolidated non-structured soils, strain weakening on shearing is affected by
the potential for strain weakening from fully softened to residual strength associated
with particle orientation.
Skempton (1985), amongst others, showed that clay content (percent finer than 2
micron) and plasticity index influence the potential for strain weakening due to particle
orientation (Figure 5.24 and Figure 5.25). Their findings indicate that strain weakening
due to particle orientation is significant for high clay content and high plasticity soils,
and is not significant for sandy soils and low to medium plasticity fine-grained soils
(clays and silts). These figures are presented merely to highlight the effects of clay
content and plasticity index, and if used should be done so with caution in consideration
of other factors affecting the residual strength of clay soils including; effective stress,
mineralogy, soil structure, pore water chemistry and rate of displacement, as
summarised by Fell et al (2000).

Figure 5.24: Residual drained strength from field case studies in clay soils in the UK
and ring shear tests on sand, kaolin and bentonite (Skempton 1985).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.52

Figure 5.25: Ring shear tests on normally consolidated sand-bentonite mixtures


(Skempton 1985, after Lupini et al 1981).

The post failure deformation behaviour of the case studies of undrained failure during
construction within wet placed earthfills provide a good example of the effect of
material type and potential for strain weakening due to particle orientation on the post
failure deformation behaviour. In the case studies considered the earthfill zones within
which failure occurred were placed well wet of optimum moisture content, and the
earthfill is characterised by low undrained shear strength, near normally consolidated
conditions under small effective normal stresses and being non-structured. The
deformation behaviour of these case studies is summarised as follows:
• Punchina cofferdam (Figure 5.26), an earthfill embankment, was constructed of silty
sands derived from deeply weathered quartz diorite placed on average 4.1% wet of
Standard optimum. The earthfill was considered as not strain weakening on
shearing due to particle orientation. The deformation of the downstream shoulder
(Figure 5.27a), during the latter stages of construction, only occurred whilst
construction was in progress and stopped shortly after cessation of construction.
This would indicate that the deformation of the unstable slope was controlled by fill
placement. Villegas (1982), by limit equilibrium analysis, demonstrated that the
factor of safety of the downstream shoulder was close to unity during the latter
stages of construction when construction rates were high (greater than 0.5 to 1
m/day). The slide also deformed as a relatively intact unit as indicated by the
reasonably uniform velocity at the different surface monitoring points (Figure
5.27b). A similar pattern of deformation behaviour was observed at Troneras dam,
an earthfill embankment constructed of wet placed medium plasticity sandy silts.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.53

• The failure during construction of Waghad dam in 1884 contrasts that at Punchina
dam. The homogeneous embankment was constructed of high plasticity, high fines
content (63 to 93% finer than 75 micron) “black cotton” soils placed without formal
compaction and at moisture contents well wet of Standard optimum (Nagarkar et al
1981). Post failure, the failed slope deformed a total distance of approximately
15 m at a velocity estimated to be in the moderate range. It is considered that strain
weakening in undrained loading contributed significantly to the amount and velocity
of the post failure deformation due to a likely large reduction in strength between
fully softened and residual strength.
• The failure during construction of Muirhead dam in 1941 (Figure 5.28) falls
somewhere between that of Punchina and Waghad dams. The puddle core earthfill
embankment was constructed of medium plasticity clays for the core and a mix of
medium plasticity clays, gravels and sands (Boulder clay) placed on the wet side of
Standard optimum for the shoulder fill. The post failure deformation behaviour
(Figure 5.29) showed continuing deformations at a decreasing velocity for some 10
to 20 days after suspension of fill placement. Placement of an additional earthfill
layer in early November 1941 resulted in an immediate resumption of deformation
that eventually totalled more than 60% of the thickness of the placed layer. The
total estimated lateral displacement of the upstream slope was in the order of 2.1 m.
It is considered that the deformation is essentially driven by fill placement as
indicated by the decreasing rate of velocity on cessation of filling, but undrained
strain weakening of the medium plasticity clays may have been a factor given the
length of time over which the deformation occurred after cessation of filling.

In the case of Waco dam and Dam FC21 failure through the foundation was along pre-
sheared high plasticity bentonite seams, the strength of which would have been at or
close to the residual strength. For these cases therefore, strain weakening due to particle
orientation was not a factor in the post failure deformation behaviour.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.54

Figure 5.26: Section through Punchina cofferdam (adapted from Villegas 1982).

(a) 2000 760

Fill Elevation (RL, m).


Deformation (mm).

1500 750

1000 740

500 SMP M1 SMP M2 730

SMP M3 Fill Elevation

0 720
0 5 10 15 20 25
Time (days since deformation observed)

(b) 400 760


Fill Elevation (RL, m).
Velocity (mm/day) .

300 SMP M1 750


SMP M2
SMP M3
200 Fill Elevation 740

100 730

0 720
0 5 10 15 20 25
Time (days since deformation observed)
Note: The SMP locations are shown in Figure 5.26.

Figure 5.27: Punchina cofferdam, (a) deformation and (b) velocity of the downstream
slope during construction (adapted from Villegas (1982)).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.55

Figure 5.28: Section through the failed region of Muirhead dam (Banks 1948)

2500 100
19/9/1941 - lateral displacement estimated at 1450 mm,
placement suspended.
First signs of deformation noted on 16/9/1941 as
displacement on stone pitching.
80

Velocity (mm/day) .
Deformation (mm) .

2000

60
Displacement
1500 Velocity
1 to 3 Nov. 1941 - 40
placement of 450
mm layer. Resulted
1000 in an increase in the
20
rate of deformation.

500 0
0 10 20 30 40 50 60 70 80 90
Time (days from 1st Sept. 1941)
Figure 5.29: Muirhead dam, post failure deformation and velocity of the failure in the upstream shoulder during construction.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.56

5.6.3.3 Failures Where Progressive Strain Weakening is Significant


The definition of significant progressive strain weakening is where on suspension of fill
placement the deformation continues to accelerate to a failure or collapse condition.
The mechanism of post failure deformation therefore involves brittleness of the slide
mass, which could be associated with strain weakening of the materials along the
surface of rupture, internal brittleness in the slide mechanics (e.g. internal shearing in
formation of a graben type failure or toe buttressing) or release from restraint on the
lateral margins (e.g. formation of a rupture surface on the margins through higher
strength earthfill zones). These types of failures are considered separately from those
where progressive strain weakening is not considered significant, i.e. cases where the
out of balance driving forces and deformation of the slide mass are largely controlled by
additional placement of fill, and on suspension of filling the deformation of the slide
mass decelerates.
Of the twenty-two case studies of failure during construction, progressive strain
weakening is considered significant in eight; Waghad, Acu, Waco, Carsington, Marshall
Creek, Chingford, Lafayette and possibly North Ridge dams. In these eight cases the
deformation behaviour either indicated acceleration of the slide mass after cessation of
filling or this was inferred from description of the failure or from the post failure travel
distance of the slide mass. The total post failure deformation of the slide mass (Figure
5.22) and the peak post failure velocity (Figure 5.23) for these cases was generally
greater than for the slides where progressive strain weakening was considered not to be
significant.
The deformation behaviour for several of the case studies is discussed in greater
detail below, but several general observations are worth highlighting:
• The slides remained virtually intact on sliding. Some break up of the slide mass was
observed at the back scarp and toe regions, as well as some break up in the mid
region of the slide. In most cases the compound shaped surface of rupture
contributed to the observed break up in the mid slope to head regions of the slide. In
several cases formation of multiple back scarps and possibly retrogressive instability
of the back scarp contributed to the break up.
• Total post failure deformations (Figure 5.22) ranged from 1.5 m (North Ridge dam)
to 25 m (Acu dam). The travel distance for most of these slides was generally
greater than 5 to 10 m.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.57

• Peak post failure velocities (Figure 5.23) were generally in the moderate range. At
Acu and Marshall Creek dams the peak post failure velocity was in the rapid range,
reflecting a high degree of brittleness in the post failure mechanics and greater
kinetic energy of the slide mass.
• Acceleration of the slide mass to peak velocity, once the failure was recognised,
ranged from about 0.5 hour (Acu dam) to 35 days (Waco dam). At Acu and
Marshall Creek dams peak velocities were reached within several hours indicating
relatively rapid development of brittleness of the slide mass. At Waco dam (Figure
5.30) and Carsington dam (Figure 5.31c) the slide mass accelerated relatively slowly
for a number days to weeks followed by a relatively abrupt increase in velocity,
possibly due to release on the lateral margins.

Of the eight cases, the factors attributed to progressive strain weakening of the slide
mass were:
• Strain weakening due to particle orientation of high plasticity, near normally
consolidated clays in either the earthfill or foundation (6 of 8 cases; exclusions are
Marshall Creek and Waco dam). At Waco the high plasticity bentonite seams in the
foundation were pre-sheared and therefore, close to the residual strength with
limited likelihood of further strain weakening. At Carsington dam (refer Section
5.5.1.3) progressive failure due to strain weakening in the over-consolidated
foundation was of greater significance than that due to particle orientation in the wet
placed core.
• Strain weakening and progressive failure within over-consolidated high plasticity
clays (Carsington dam).
• Strain weakening due to contraction on shear and static liquefaction. At Marshall
Creek dam (refer below) the rapid development of a failure condition was possibly
associated with static liquefaction within the low to high plasticity loessic clays in
the foundation.
• Internal brittleness in the slide mechanics. This was considered a factor in the
failure of Acu dam (see below) due to the embankment zoning geometry, and at
Waco dam due to passive resistance in the toe region of the slide (see below).
• Strain weakening due to release or reduction of restraining forces on the lateral
margins. Two factors are considered significant: the lateral persistence of weak
layers in either the foundation or earthfill (Carsington and possibly Lafayette dam);
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.58

and the formation of a surface of rupture on the margins through embankment zones
of higher strength (Acu, Waco and Carsington dams). In either case, strength
reduction due to strain weakening is significant. Either strain weakening within the
weaker section of the embankment occurs and increases the stresses on the margins
such that failure occurs, or else strain weakening on the lateral margins occurs. In
all eight cases progressive strain weakening within the weaker section of the
embankment was considered a factor resulting in increased shear stresses to be
supported on the lateral margins.

Further details on the failure and deformation behaviour at Waco, Carsington, Acu,
Marshall Creek and Chingford dams are discussed below.

(a) Waco Dam


The failure at Waco dam (Beene 1967; Stroman et al 1984) occurred in 1961. The
zoned earthfill embankment (Figure 5.13b) consisted of a central core zone of rolled,
well-compacted clayey soils and shoulders of well-compacted earthfills. The surface of
rupture within the foundation was along a pre-sheared bentonitic clay seam in the
Pepper Shale formation, bounded at the margins by near vertical fault zones. Along
both the basal and lateral slide planes (in the foundation) the strength at failure was at
residual strength due to pre-shearing, and therefore strain weakening on shearing at the
time of the slide did not occur. The near vertical fault zones were oriented slightly
offset from perpendicular to the embankment centreline and to each other (the width
between the fault zones increased in the downstream direction), which was a significant
factor in the downstream direction of failure.
The slide was first noticed on 4 October 1961 when cracks were observed on the
downstream slope. The cracks consisted of a longitudinal crack (parallel to the
embankment centreline) about one third of the distance down the slope from the
centreline, and small diagonal cracks at the downstream toe located approximately
above the two fault zones at the margins of the eventual slide. Beene (1967) thought
that the failure may have initiated some two weeks prior and was obscured from earlier
observation by embankment construction and the presence of riprap on the upstream
slope.
The monitored deformation of the crest after 4 October 1961, Figure 5.30, shows the
slow development of the slide. The velocity of deformation increased relatively quickly
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.59

after about 33 days (from when the slide was thought to have initiated) and reached a
maximum at about 36 days before slowing significantly.
The significant factors affecting the slide deformation are thought to include:
• Passive resistance at the toe of the slide. The toe of the slide was located some 140
m downstream from the downstream toe of the embankment, and the basal surface
of rupture some 14 m below ground surface level.
• No strain weakening along the basal surface of rupture.
• Overall strain weakening of the slide mass, as evidenced by the gradually increasing
velocity of deformation, was considered to be associated with strain weakening on
the margins and back-scarp in the earthfill, and weakening in the toe region.
• The large increase in velocity of deformation from Day 33 is thought to be
associated with release on the lateral margins and reduction in passive resistance in
the toe region. The pulling away of the slide mass from the north fault is also likely
to have been a contributing factor.

8000 2000

7000 Settlement at crest 1750


Crest Settlement (mm) .

.
Velocity
6000 1500

Velocity (mm/day)
5000 1250

4000 1000

3000 750

2000 500

1000 250

0 0
0 10 20 30 40 50 60
Time (days since start of deformation on 17/9/61)
Figure 5.30: Crest settlement monitoring of the slide at Waco dam (adapted from
Stroman et al 1984).

(b) Carsington Dam


Section 5.5.1.3 briefly discusses the earthfill materials and placement methods used, and
the mechanics of progressive failure leading up to the failure in the upstream shoulder at
Carsington dam. The slide occurred over a period of about 5 to 6 days.
The surface monitoring at the upstream toe, when analysed in hindsight, indicated a
small (49 mm) but significant increase in the lateral displacement over the period from
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.60

30th May to 1st June 1984 (at chainages 675 and 750 m only) in the last stages of fill
placement prior to the failure (Figure 5.31a). Following a weekend of wet weather it
was noted on Monday 4th June that an additional 75 mm of displacement had occurred
in the toe region and a tension crack was observed on the embankment crest between
chainages 670 and 790 m. Over the next 30 hours the width and velocity of the slide
increased significantly (Figure 5.31c) culminating in a large deformation (10.4 m)
overnight on the 5th June (Figure 5.31b). By the 7th June the deformation had virtually
ceased.
There are several important observations associated with the pre and post failure
deformation of the slide and the slide mechanics, they are:
• During the winter shutdown period starting in October of 1983, the IVM (internal
vertical settlement gauge) within the core at chainage 850 m indicated an increase in
vertical compression of the fill from 7 to 9% before it became blocked. This
location was coincident with the eventual surface of rupture.
• Over the period 24th May to 4th June 1984 the IVM within the core at chainage
705 m showed a large (2.5 to 5%) increase in vertical compression over the depth
range 10 to 15 m. This equates to a vertical deformation in this zone of some 225
mm for placement of less than 1 m thickness of fill.
• The surface monitoring on the lower upstream slope (Figure 5.31a) gave little
warning of the impending failure, only some 2 to 3 days. This is not unexpected
given the mechanics of failure (Section 5.5.1.3) and findings from the Potts et al
(1990) analysis indicating the surface of rupture had not formed in the toe region
prior to the failure.
• Over the course of the slide the width of the failed section increased from about 120
m on the 4th June to 500 m by the 6th June.

Skempton and Vaughan (1993) commented that significant restraining forces were
exerted by the as yet unfailed portions of the embankment on either side of the failed
portion of the embankment (approximately from chainage 660 to 800 m). Toward the
gully region significant restraint was afforded due to the lack of lateral persistence of
the “yellow clay” foundation (from chainages 800 to greater than 900 m the foundation
was on weathered mudstone). The effect of the lateral restraint is likely to have been
significant in suppressing the amount of deformation within the unstable portion of the
slope prior to and in the early stages of failure.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.61

(a) 300 202


Ch 750m (SMP at upstream toe) Marked Increase in velocity of
Ch 875m (SMP at upstream toe) displacement on 31 May to 1 June 1984
Deformation at toe (mm) . 250 at both Chainage 750 m and 675 m. 200
Fill Elevation (RL, m)

Fill Elevation (RL, m)


200 198

150 196

100 194

50 29/5/84 (Day 289) 192

1/6/84 (Day 292)


0 190
0 50 100 150 200 250 300
Time (days since 15th August 1983)

(b) 16000 202

14000 Ch 750m (SMP at upstream toe)


200
Deformation at toe (mm) .

Ch 875m (SMP at upstream toe)


12000

Fill Elevation (RL, m)


Fill Elevation (RL, m)
198
10000
Large acceleration of slide
mass overnight on 5/6/84
8000 196

6000
194
4000
192
2000

0 190
286 288 290 292 294 296 298
Time (days since 15th August 1983)

(c) 20000
Ch 750m (SMP at upstream toe)
Velocity at toe (mm/day) .

15000 Ch 875m (SMP at upstream toe)


Note: Day 291 is 1 June

10000

5000

0
291 292 293 294 295 296 297 298
Time (days from 15th August 1983)

Figure 5.31: Carsington dam, deformation monitoring of upstream shoulder (a) leading
up to failure, (b) during failure (Skempton and Vaughan 1993), and (c) the velocity of
the slide mass.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.62

A significant factor in the observed increase in velocity of the slide mass (Figure 5.31c)
over time was strain weakening on shearing within the failed portion of the
embankment, thereby increasing the shear stresses on the lateral margins. Once the
shear capacity of the lateral margins was reached the slide mass accelerated at a much
faster rate and the width of the failure spread from about 120 m to 500 m.

(c) Acu Dam


In the case of Acu Dam (Figure 5.32) the surface of rupture passed through a wet
placed, high plasticity clay layer at the base of the upstream shoulder and extended back
through the central high plasticity clay core (initially placed wet of optimum then
adjusted to close to optimum in the mid to upper portion of the core). The failure, as
described by Penman (1985), was initially observed as a tension crack at crest level on
the downstream edge of the core zone. This was followed by sinking of the core and
upstream shoulder, and a massive pushing outward of the lower section some 25 m in
distance over a width of some 600 m. The deformation at and post failure took place
over a period of about 30 minutes (average velocity of 50 m/hour or 1.2 x 106 mm/day)
putting the failure into the rapid velocity category (Figure 5.23).
The large post failure travel distance of 25 m and rapid velocity of the slide mass are
indicative of a relatively high degree of brittleness of the overall slide mechanism.
Contributing factors to the high degree of brittleness are considered to include; a
buttressing effect of the sandy and gravelly earthfill in the toe region of the slide,
internal shearing in development of the compound slide mechanism, undrained strain
weakening of the high plasticity wet clays and development of the back-scarp through
partially saturated earthfill. The extensive lateral persistence of the layer of clay
earthfill under the upstream slope and the high shear strength of the shoulder fill
(compacted clayey sands and gravels) are considered to have contributed to the broad
width (600 m) of the slide.

Figure 5.32: Section through failed region of Acu dam (Penman 1986)
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.63

(d) Marshall Creek Dam


The slide at Marshall Creek dam (Sherard et al 1963; ENR 1937a, 1937c, 1938a and
1938b) was located within the relatively broad gully region of the foundation. The
failure occurred within several hours of the observation of cracking on the embankment
crest. The majority of the post failure deformation (about 15 m) occurred at a rapid
velocity over a period of 15 to 20 minutes. The relatively large post failure travel
distance and rapid velocity of the slide mass are indicative of a relatively high degree of
brittleness of the overall slide mechanism. Significant strain weakening associated with
static flow liquefaction of the saturated loess foundation is possibly the reason for the
high degree of brittleness.

(e) Chingford Dam


The failure of Chingford dam (Cooling and Golder 1942) consisted of a compound type
failure (Figure 5.33) with basal translational component within the very soft, highly
plastic clay foundation and rotational back-scarp through the soft to very soft, highly
plastic core and select earthfill zones. The relatively high rate of construction
(approximately 100 mm/day) of this relatively low height (7.9 m height at failure)
puddle core earthfill dam was considered to be a significant factor in the undrained
slope failure.
Breakage of a water pipe at the toe of the slope several days prior to the slide was an
indication of the impending slide. The slide travelled a distance of 4.3 m at the toe at an
estimated peak post failure velocity in the moderate range. Undrained strain weakening
of the near normally consolidated highly plastic clay foundation and embankment
materials is considered to have been the major factor contributing to strain weakening of
the slide mass.

Figure 5.33: Cross section of the slide during construction at Chingford dam (Cooling
and Golder 1942)
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.64

5.6.3.4 Failures Where Progressive Strain Weakening was Not Significant


For failures where progressive strain weakening is not considered significant, the
deformation of the slide mass is largely controlled by the placement of fill, which
provides the out of balance driving forces for the deformation to occur. Deceleration of
the slide mass is generally observed for these cases on suspension of filling, such as
observed at Muirhead dam (Figure 5.29) and Punchina cofferdam (Figure 5.27).
In at least ten of the twenty-two case studies of failure during construction,
progressive strain weakening of the slide mass was considered as not significant (Otter
Brook, Truscott, Skiatook, Troneras, Punchina, West Branch, Dam FC21, Muirhead,
Hollowell and Galisteo dams). The failures at Lake Shelbyville and Scout Reservation
dams were considered marginal (discussed below) and for Park Reservoir insufficient
information was available on the deformation behaviour and material types for
evaluation of the influence of progressive strain weakening of the slide mass. The
material types of these case studies included:
• Wet placed silty sand and low to medium plasticity sandy silt earthfills (Punchina
cofferdam and Troneras dam).
• Wet placed earthfill layers or non-structured low strength foundations consisting of
low to medium plasticity clays (Otter Brook, Truscott, Skiatook, West Branch,
Muirhead and Galisteo dams). For these cases the materials within which the basal
portion of the surface of rupture was located were considered to be near normally
consolidated.
• Deformation along pre-sheared high plasticity bentonitic seams in the foundation
(Dam FC21).

The failure at Hollowell dam (Kennard 1955), a puddle core earthfill embankment,
is an exception to the general material types. The basal portion of the surface of rupture
was located within medium to high plasticity clays of soft to very soft strength
consistency in the foundation, with a back-scarp through the puddle and select high
plasticity core zone of the embankment. Progressive strain weakening was considered
as not significant for the overall slide even though strain weakening on shearing of the
high plasticity soils was likely to have occurred. This is because of the limited lateral
extent of the weak foundation (30 to 40 m width) and the capacity of the lateral margins
to accommodate the increased shear stresses associated with the strain weakening in the
narrow failure section. The deformation of the slide was limited to less than 500 mm
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.65

and was estimated to have occurred at a slow velocity. The overall width of the
embankment affected by the slide was about 200 m, but the bulk of the deformation was
confined to the embankment section over the soft to very soft clay foundation layer.
For the case studies where progressive strain weakening was not considered
significant, total deformations of the slide mass were limited to 0.5 m to 2 m (Figure
5.22) and peak velocities generally to the slow class (Figure 5.23). Several aspects of
the deformation behaviour for these case studies is important in the context of
identifying potential instability and undertaking remedial measures, they are:
• For the case studies controlled by wet placed earthfill layers or low undrained
strength foundation conditions, the failures (including the cases where progressive
strain weakening was considered significant) were usually recognised by cracking
on the embankment crest or slopes, and in some cases by cracking in the toe region
of the slide, movement of the outer slope or spreading in conduits. For these cases
the deformation of the outer slopes was significant when the slide was first
recognised. This was considered to be because the critically stressed weak layer
formed the basal portion of the surface of rupture within which plastic deformations
had occurred. Therefore, deformation monitoring for identification of potential
instability in these case studies is best undertaken in the region from the toe to the
mid slope of the embankment.
• The lengthy time for development of the surface of rupture. On recognition of a
potentially unstable slope condition remedial measures can be (and were)
undertaken to improve the factor of safety of the embankment slope. The remedial
measures generally took the form of addition of stabilising berms or changes in the
embankment design, or in some cases adjustment in the construction program to
allow construction pore water pressures to dissipate.

Further details on the failure and deformation behaviour of several of the marginal
case studies (Lake Shelbyville, Scout Reservation and Dam FC10) are discussed in the
following sections.

(a) Lake Shelbyville dam


In the case of Lake Shelbyville dam, an earthfill embankment constructed using sandy
clays of low to medium plasticity, a slide occurred in the closure section. The slide was
not recognised until after the end of construction. Wet placed earthfill layers formed the
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.66

basal surface of rupture. The inclinometer records and development of back-scarps at


the embankment crest confirmed the full development of the surface of rupture in cross
section, but the slide may not have fully developed on the lateral margins.
The published deformation records (Figure 5.34) show the “slow” velocity of post
failure deformation (generally less than 0.5 mm/day) that persisted over a period of up
to 2 years after the end of construction. During this time construction pore water
pressures were still present.
Prior to completion of embankment construction the reservoir impounded water, and
it was on drawdown, prior to proposed remedial works, that acceleration of the slide
mass was observed. This suggests that the water load on the upstream slope acted as
support for the unstable portion of the upstream slope. The persistence of deformation
for several years after the end of construction, during which time pore water pressures
are likely to have dissipated to some degree, is considered to probably reflect some
degree of strain weakening of the overall slide mechanism.

(b) Scout Reservation dam


The slide at Scout Reservation dam (Figure 5.13a), an earthfill embankment constructed
using sandy clays of low to medium plasticity, was not recognised until after the end of
construction. Inclinometer records and development of back-scarps at the embankment
crest indicated that a surface of rupture had fully developed in cross section.
The published deformation records (Figure 5.35) show the “slow” post failure
velocity of the slide mass (less than 10 mm/day) that persisted over a period of up to 2
years after the end of construction, during which time construction pore water pressures
were still present. This persistence of deformation over several years, during which
time pore water pressures are likely to have dissipated to some degree, is considered to
reflect strain weakening of the overall slide mechanism. Deformations that occurred
more than 2 years after the end of construction were triggered by rainfall, most likely
due to water infilled cracks at the crest of the embankment.

(c) Dam FC10


Dam FC10 is a zoned earth and rockfill embankment with broad central clay core of
high plasticity and high clay content (35% passing 2 micron) supported by thin outer
random fill and rockfill zones on the upstream shoulder. Monitoring during
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.67

construction indicated the deformation behaviour of the upstream shoulder was


“abnormal”, and the slope to potentially be at a limiting stability condition.

500 2.5
SMP M3 (inside failure zone)
SMP M1 (outside failure zone)
400 SMP M3, velocity 2

Velocity (mm/day)
Deformation (mm)

Acceleration due to drawdown


300 prior to start of remedial works 1.5

200 1

100 0.5

0 0
0 200 400 600 800 1000 1200 1400
Time (days since end of construction)

Figure 5.34: Deformation and velocity of the slide at Lake Shelbyville dam (adapted
from Humphrey and Leonards (1986, 1988)).

2000 10
.
Crest settlement (mm)

.
1500 7.5

Velocity (mm/day)
velocity
crest settlement
1000 5

500 2.5

0 0
0 200 400 600 800 1000 1200 1400
Time (days since end of construction)

Figure 5.35: Post failure deformation and velocity of slide at Scout Reservation dam
(adapted from Mann and Snow (1992)).

The total measured lateral displacement of the upstream slope during construction was
390 mm, which was significantly greater than for the downstream slope. Of this, some
230 to 260 mm occurred during placement of the final 2.5 m (13% of the total height).
Shortly after the end of construction the deformations were observed to accelerate
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.68

reaching a total of some 2.8 m over a period of many days. This acceleration may have
been triggered by a reduction in the reservoir level.
Strain weakening of the wet placed, near normally consolidated high plasticity broad
core zone at Dam FC10 is likely to have occurred. But, the overall effect of the strain
weakening was not greatly significant, as indicated by the deformation behaviour of the
slide. The zoned earth and rockfill geometry is likely to have been significant in the
limited total deformation of 2.8 m measured. Cases such as Dam FC10 are considered
marginal in that strain weakening on shearing is a factor in the mechanics of post failure
deformation and under different design circumstances post failure strain weakening may
have been significant.

5.6.4 FAILURES DURING DRAWDOWN – POST FAILURE DEFORMATION BEHAVIOUR

Of the nineteen failures in embankment dams that occurred during drawdown, sixteen
were within the embankment only and the remaining three through the embankment and
foundation. The embankment types (Table 5.7) consisted predominantly of earthfill
embankments; homogeneous earthfill (6 possibly 9 cases), earthfill with filters (2
cases), and earthfill with rock toe (3 cases). For the two failures in embankments with
upstream rockfill zones, the surface of rupture preferentially passed through the
foundation and not the outer rockfill zones.
Table 5.14 presents a summary of the timing of the failure, embankment type and
geometry, material properties, drawdown causing failure and comments on the slide
itself. Further details are given in Section 1 of Appendix D. Table 5.15 presents a
summary of the post failure deformation behaviour (total deformation and estimated
peak velocity), slope failure geometry and material properties. Plots of the estimated
total post-failure deformation and estimated peak post failure velocity are presented in
Figure 5.36.
The deformation records of the case studies are limited. In most cases information
was only available on the total deformation and in some cases this was estimated off
sections provided in the literature. The peak velocities have been estimated from the
monitoring data where possible (4 cases) or a velocity range has been estimated based
on the amount of deformation and reference to times over which the deformation
occurred. These estimates are broad range.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.69

Table 5.14: Summary of the case studies of slides in the upstream shoulder of embankment dams triggered by drawdown
Embankment Details Material Properties (through which
Years
Year of surface of rupture passed) Drawdown
Name since Comments
Failure/s Type Hgt Slope Embankment *1 Foundation *1 (causing failure)
Const d
(m) (o)
Old Eildon 1927 & 2 to 4 Conc. C- 29 36 Puddle clay zone (CL – CL to CH – alluvial First two drawdowns. 1927 - 32.5 Deep seated compound slide passing
1929 wall R’fill medium plasticity) upstream of clays of firm to stiff m at 200 mm/day (increasing to through the puddle core and
(10,0,0) corewall. strength consistency 500), and 1929 - 15.5 m at 100 foundation. Back-scarp at corewall
Dumped rockfill shoulder. mm/day (increasing to 230). (centre of crest). Crest settled 14 m
1929.
Wassy 1883 1 Homog 16.5 33.7 CL –silty clays, medium (failure through First drawdown, very rapid (10 m Rotational slide through the outer
(0,0,0) plasticity. No formal embankment only) drawdown at 400 to 500 mm/day) earthfill zone of the steep upstream
compaction shoulder. Back-scarp at upstream
edge of crest. Brick facing on
upstream slope.
Mondely 1981 1 E’fill - 24 18.4 CL – low to medium plasticity (failure through First drawdown. Re-activation of Relatively shallow rotational slide in
filters clays placed wet of OMC. embankment only) deformation on second the upper part of the upstream slope.
(1,1,1) Rolled, well compacted. drawdown. Limited deformation. High pore
water pressures from construction
still present in the earthfill.
Malko unk 1 Zoned 41 unk CL (?) – sandy clay in outer (failure through First drawdown. No information on the slide.
Sharkovo E’fill earthfill upstream shoulder. embankment only)
(3,0,1)
Forsythe 1921 1 Homog 19.8 26.6 SM/ML – silty sands and sandy (failure through First drawdown, very rapid (4.6 Rotational slide through the outer
(0,0,0) silts of low plasticity. No embankment only) m drawdown in several hours, earthfill zone of the steep upstream
formal compaction. rate ˜ 2.3 m/hour) slope. Back-scarp at the centre of the
crest.
San Luis 1981 14 Zoned 56 15.8 CL – sandy clays, medium CH – colluvial sandy Largest and quickest drawdown Deep seated compound slide through
E&R plasticity. Rolled, well clays, low in embankment history in 1981 the broad central core, foundation and
(4,2,2) compacted. Broad central core. permeability, (55 m at max. 50 day rate of 625 random fill zones. The slide travelled
fissured. mm/day) some 17 m.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.70

Table 5.14 (cont.): Summary of the case studies of slides in the upstream shoulder of embankment dams triggered by drawdown (Sheet 2 of 3).
Embankment Details Material Properties (through which
Years
Year of surface of rupture passed) Drawdown
Name since Comments
Failure/s Type Hgt Slope Embankment *1 Foundation *1 (causing failure)
Const d
(m) (o)
Charmes 1909 3 Homog 16.8 33.7 CL – silty sandy clays of (failure through Largest drawdown in Rotational slide through the outer
(0,0,0) medium plasticity. Rolled, embankment only) embankment history. earthfill zone of the steep upstream
limited compaction. slope. Back-scarp at centre of crest.
Slow post failure deformation.
Enista 1982 13 E’fill - 32.2 18.4 CL/CH (?) – alluvial silty clays. (failure through Drawdown to lowest historical Incipient slide. 400 mm wide cracks
filters Rolled, well compacted. embankment only) level. in upstream slope, no vertical
(1,0,1) displacement across cracks.
Dam FD2 1993 32 Homog 28.7 26.6 CL – sandy clays, low (failure through Likely failure during 1993 Relatively shallow (5.5 m depth)
(0,0,0) plasticity. Rolled, well embankment only) drawdown, largest and quickest compound slide. Reactivation of
compacted. in history of operation (16.7 m at deformation during 1994 drawdown.
335 to 400 mm/day).
Mount 1928 18 Homog 23.2 33.7 ML – sandy silts. Poor (failure through 13.4 m drawdown over several Rotational slide in the steep upstream
Pisgah (0,0,0) construction practice. embankment only) months (average 220 mm/day). slope. Back-scarp at downstream
Drawdown very rapid over edge of crest. Rapid deformation.
several days prior to failure. Concrete facing on upstream slope.
Pilarcitos 1969 103 Puddle 29 21.8 GC – clayey sandy gravel of (failure through Very rapid drawdown (7.6 m at Rotational slide in the upstream
E’fill low plasticity (upstream embankment only) 540 mm/day), possibly the most slope. Limited deformation (˜ 1 m)
(8,0,0) shoulder zone). No formal rapid historical drawdown.
compaction.
Sampna 1961 & 4 and 7 Zoned 17.8 26.6 CL/CH (?) – clayey earthfill (failure through 1964 drawdown to lowest Rotational slide in outer earthfill zone
Tank 1964 E’fill used in embankment shoulders. embankment only) historical level. Drawn down 9.1 of the embankment. Back-scarp at
(3,0,0) m over 8 months (average 30 to upstream edge of crest. Rapid post
60 mm/day) failure deformation.
Waghad 1919 34 E’fill – 26.9 14 CH/MH – alluvial clays and (failure through Large drawdown (up to 14 m), no Deep seated rotational slide that
rock toe silts of high plasticity. Placed embankment only) indication of rate or if abnormally extended back to the downstream
(2,0,0) well wet of OMC, no formal large. edge of the crest. Large post failure
compaction. deformation of the slide mass.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.71

Table 5.14 (cont.): Summary of the case studies of slides in the upstream shoulder of embankment dams triggered by drawdown (Sheet 3 of 3).
Embankment Details Material Properties (through which
Years
Year of surface of rupture passed) Drawdown
Name since Comments
Failure/s Type Hgt Slope Embankment *1 Foundation *1 (causing failure)
Const d
(m) (o)
Dam FD3 1931 20 E’fill – 27.4 26.6 CL/CH – silty clays of medium (failure through Large but not unprecedented Shallow rotational slide in the steep
rock toe to high plasticity. Rolled, embankment only) drawdown (10.3 m drawdown at upstream slope. Possible rapid post
(2,0,0) limited compaction. 110 mm/day) failure deformation at the toe of the
slide. Concrete faced upstream
shoulder.
Bear Gulch 1936, 1942 40 to 48 Homog 19.4 18.4 CL – sandy clays of medium Soft and wet clay Similar drawdowns in 1936, 1942 Compound slide in the upstream
& 1944 (6 to 14 (0,0,0) ? plasticity. No formal layer in foundation and 1944 causing deformation (8 slope through the embankment and
after compaction. (interbedded m drawdown at average 90 foundation. Limited post failure
raising) sandstones and clay mm/day) deformation.
shales)
Mechka 1981 11 Homog 32 18.4 (?) CL (?) – sandy silty clays (failure through Routine drawdown. No indication of location or size of
(unk) ? embankment only) slide, or amount of deformation.
Sushitsa 1981 11 Homog 32.5 18.4 (?) CL (?) – sandy silty clays (failure through Routine drawdown. No indication of location or size of
(unk) ? embankment only) slide, or amount of deformation.
Telish 1982 14 Homog 29.5 18.4 CL (?) – sandy silty clays (failure through Routine drawdown. Rotational slide in the upstream
(0,0,0) ? embankment only) shoulder. Limited post failure
deformation.
Drenovets 1984 16 E’fill – 29 18.4 CL (?) – sandy silty clays (failure through Routine drawdown. No indication of location or size of
rock toe embankment only) the slide, or amount of post failure
(2,0,0) deformation.
Notes: Hgt = height Const d = construction unk = unknown OMC = Standard optimum moisture content Homog = homogeneous earthfill embankment
E’fill – filters = earthfill embankment with filters E’fill – rock toe = earthfill embankment with rock toe Zoned E’fill = zoned earthfill embankment
Puddle E’fill = puddle core earthfill embankment Zoned E&R = zoned earth and rockfill embankment
Conc. C-wall R’fill = concrete core-wall rockfill embankment
*1 Soil classification to the Australian soil classification system (AS 1726-1993)
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.72

Table 5.15: Post failure deformation behaviour and slope/material properties of slides in
embankment dams during drawdown.
Dam Name Slide Estimated Peak Slope Upstream ASCS
Deformation Velocity Failure Slope Angle Classn
(m) (mm/day) Geometry*1 (degrees) (*2)
San Luis dam 15 (crest), 1000 to 2000 Type 2 15.8 CI (emb), CH
17 (toe) (fndn)
Waghad 4.8 (crest), 4000 to 40000 (est.) Type 2 18.4 CH/MH
10 (toe)
Sampna Tank 11.9 (crest) 50000 to 200000 Type 1 (?) 26.6 CI (upstream
(est.) earthfill)
Dam FD3 1.7 (crest), 50000 to 500000 Type 1 26.6 CI
6 to 9 (toe) (est.)
Old Eildon 14 (crest), 5000 Type 2 36 CI (puddle
17 (toe) core and
fndn)
Mount Pisgah 7 (crest), 10000 to 400000 Type 2 (?) 33.7 ML
9 (toe) (est.)
Forsythe 4 (crest), 100000 to 1000000 Type 1 (?) 26.6 ML/SM
22 ? (toe) (est.)
Bear Gulch 0.32 (crest) 10 to 200 (est.) Type 2 18.4 CI
Mondely 0.4 (crest) 10 to 30 Type 2 (?) 18.4 CI
Enista 0.4 (crest) 50 to 400 Type 2 (?) 18.4 CL/CH (?)
Dam FD2 0.42 27 Type 2 26.6 CL
Telish 1 (crest) 100 to 500 (est.) Type 2 (?) 18.4 CL (?)
Pilarcitos 1 est. (crest) 200 to 500 (est.) ? 21.8 GC
Notes: crest = crest of back-scarp toe = toe of slide est. = estimated
emb = embankment fndn = foundation
*1 refers to the location of the slide in the slope, see Section 1.3 for definitions.
*2 = Australian soil classification system, CI refers to clays of medium plasticity

From analysis of the deformation data, the significant factors influencing the post
failure deformation behaviour of slides in embankment dams triggered by drawdown
(and discussed further in the following sub-sections) include:
• The material type of the earthfill and foundation materials within which the surface
of rupture is located.
• The slope failure geometry.
• The slope angle of the upstream slope.
• Factors that add brittleness to the mechanics of failure and sliding at and post
failure, such as slope facing materials (e.g. concrete facing).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.73

• Factors contributing to restraint on the lateral margins including lateral persistence


of weak layers, shearing through higher strength outer earth or rockfill zones, and
changes in foundation topography.

(a) 25
CH/MH - high Clays (CL) Clays and Silts (CL or Sandy and
plasticity medium plasticity ML) of low plasticity Gravelly Soils
?
Post Failure Deformation (m) .

20

Deformation at head of back-scarp


15
Deformation at toe of slide

10

0
Dam FD3

Dam FD2
Enista

Telish
Mondely
San Luis

Bear Gulch

Pilarcitos

Forsythe
Waghad

Old Eildon

Mount Pisgah
Sampna Tank

(b) 1.E+07
CH/MH - high Clays (CL) Clays and Silts (CL or ML) Sandy and
plasticity medium plasticity of low plasticity Grav. Soils
1.E+06
Peak Velocity (mm/day) .

? Rapid
? ?
1.E+05 ?

1.E+04 ?
Moderate
1.E+03

? ?
1.E+02 ?
? Slow
1.E+01

Estimate from monitoring data Broad range estimate Very


1.E+00
Slow
Dam FD3

Dam FD2
Enista

Telish
Mondely
San Luis

Bear Gulch

Pilarcitos

Forsythe
Waghad

Old Eildon

Mount Pisgah
Sampna Tank

Figure 5.36: Slides in embankment dams during drawdown; post failure (a) deformation
and (b) estimated velocity.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.74

5.6.4.1 Material Type and Strain Weakening Potential


Material type has a significant effect on the potential for strain weakening on shearing,
leading to potentially large deformations at failure. As discussed for the failures during
construction (Sections 5.6.3.2 and 5.6.3.3), material types with high strain weakening
potential include:
• Near normally consolidated high plasticity clays or silts with a high clay fraction,
which have significant strain weakening potential due to particle orientation.
• Soil types, in a saturated or near saturated condition, that are potentially contractile
on shearing.
• Progressive failure in over-consolidated soil types (refer Section 5.2.2).

Of the case studies of failure during drawdown, material strain weakening from the
failure condition was considered significant at Waghad, San Luis, and possibly Sampna
Tank, Dam FD3 and Old Eildon dams. In these cases the surface of rupture passed
through medium to high plasticity soil types. At Waghad dam (wet placed near
normally consolidated earthfill) strain weakening due to particle orientation in the high
plasticity earthfill was considered significant. At Sampna Tank strain weakening due to
particle orientation and possibly progressive failure (mainly due to strength loss on
shearing in partially saturated soils) in the medium to high plasticity outer earthfill were
considered significant factors in the large post failure travel distance of 12 m for this
slide. At San Luis (refer below) strain weakening due to progressive failure and particle
orientation were considered significant factors in the relatively large post failure travel
distance of 15 to 17 m.
Of the remaining cases material strain weakening on shearing was not considered
greatly significant in the mechanics of post failure deformation. For several of these
cases relatively large deformations were observed at failure due to factors other than
material strain weakening, but in most cases post failure deformations were less than
about 1 m and peak post failure velocities estimated to be in the slow class for failures
in low to medium plasticity clays and silts, and in sandy and gravelly soil types.
In the discussion on the mechanics of failure (Section 5.5.2), progressive strain
weakening in undrained loading of over-consolidated earthfills was considered
significant in the mechanics of failure for a number of these cases. Athough, the limited
post failure deformation would suggest it was not significant in the mechanics of post
failure deformation. It is possible that a reduction of pore water pressures on shearing
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.75

within the shear band or along the surface of rupture retarded the post failure
deformation in several of these cases, such as at Dam FD2 (discussed below).

(a) San Luis dam


The drawdown failure at San Luis dam (Figure 5.17) consisted of a deep-seated failure
through the high plasticity slope wash foundation, broad central medium plasticity clay
core and outer random earthfill zone. As discussed in Section 5.5.2.2, progressive
failure was considered a significant factor in the mechanics of failure and would also
have been significant for the mechanics of post failure deformation.
The slide itself was not observed until 14th September 1981 when the back-scarp
was at a height of about 8 m. Photographic evidence indicated that limited
deformations had occurred by the 4th September, but they were obscured by the riprap
facing on the upstream slope. It is possible that the slide initiated in early to mid
August toward the end of the very large drawdown.
The deformation monitoring only picked up the deceleration component of the slide
movement (Figure 5.37). The records show that the post failure deformations were
ongoing for some 100 days after the end of drawdown (about 28th August) and that the
slide velocity gradually decreased over a period of 50 to 60 days. The likely period of
peak velocity is estimated to have been between the 10th to 12th September and is
broadly estimated at 1000 to 2000 mm/day. The velocity profile of the slide is therefore
likely to be similar to that at Waco dam (Figure 5.30) and Carsington dam (Figure
5.31c), where the velocity increased slowly for several to many days prior to rapid
acceleration to peak velocity followed by a period of deceleration. This type of
deformation behaviour is indicative of progressive strain weakening during post failure
deformation.
Factors contributing to the brittleness of the slide are likely to include; material
strain weakening, due to both particle orientation in high plasticity clays and strain
weakening in undrained loading of over-consolidated earthfills (Stark and Duncan
1991), and restraint on the lateral margins due to the variation in foundation topography
(Von Thun 1988).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.76

(a) 20000 170

Reservoir Level (RL, m)


16000 150
Deformation (mm)

12000 130

8000 110
30 Sept 1981, start of
61 m upstream (near head of slide)
deformation monitoring
4000 106 m upstream (mid slope of slide) 90
146 m upstream (mid slope of slide) 14 Sept 1981, slide first
reservoir recognised
0 70
-150 -100 -50 0 50 100 150
Time (days since 14 August 1981)

(b) 350 170


Likely period when
300 slide initiated

Reservoir Level (RL, m)


150
Velocity (mm/day)

250

130
200

150
110

100 61 m upstream (near head of slide)


106 m upstream (mid slope of slide) 90
50 146 m upstream (mid slope of slide)
reservoir
0 70
-150 -100 -50 0 50 100 150
Time (days since 14 August 1981)
Figure 5.37: San Luis dam; the latter stages of post failure (a) deformation and (b)
velocity of the first time slide following the 1981 drawdown (data courtesy of U.S.
Bureau of Reclamation).

(b) Dam FD2


The deformation behaviour of the reactivated slide at Dam FD2 contrasts that of the first
time slide at San Luis dam. The failure at Dam FD2 (Figure 5.38) occurred in the
homogeneous left abutment section of the embankment in earthfills comprising well-
compacted low plasticity sandy clays. The slide was thought to have initially occurred
during the historically large and rapid 1993 drawdown (Figure 5.18a). The monitored
deformation behaviour (Figure 5.39) is from reactivation of the slide during the
subsequent 1994 drawdown.
Deformation of the slide was reactivated when the reservoir level was
approximately coincident with the centre of mass of the slide. The interesting aspects of
the deformation behaviour are that the slide moved as an intact unit and the velocity was
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.77

relatively constant whilst the drawdown rate was constant. Once the rate of drawdown
slowed the velocity of the slide mass also decreased (Figure 5.39b), and continued for
some 10 to 15 days after the drawdown rate was significantly reduced. Therefore, the
out of balance forces due to the pore water pressure conditions in the upstream slope
and reducing support of the reservoir water load on the upstream face were driving the
deformation during reactivation of the slide. Progressive strain weakening is not
significant during the reactivation as indicated by the constant velocity under constant
drawdown rate and deceleration on decrease in the drawdown rate.

Figure 5.38: Section through the slide at Dam FD2.

(a) 500 1673


Reservoir Level (RL, m).
Deformation (mm) .

400 1670

300 1667

200 1664

100 1661

0 1658
0 20 40 60 80 100
Time (days since 9 May 1994)
14.6 m upstream 21.3 m upstream 26.2 m upstream 31 m upstream Reservoir
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.78

(b) 40 1673

Reservoir Level (RL, m)


Velocity (mm/day) 1670
30

1667
20
1664

10
1661

0 1658
0 20 40 60 80 100
Time (days since 9 May 1994)
14.6 m upstream 21.3 m upstream 26.2 m upstream 31 m upstream Reservoir

Figure 5.39: Deformation behaviour of the slide at Dam FD2 on reactivation during the
1994 drawdown; (a) deformation and (b) velocity.

5.6.4.2 Slope Failure Geometry and Upstream Slope Angle

The slope failure geometry (definitions given in Section 1.3 of Chapter 1) and upstream
slope angle are significant factors in the post failure deformation behaviour of the slides
during drawdown.
The failures at Dam FD3 and Forsythe dam were of Type 1 slope failure geometry.
In these cases the toe of slide mass broke up and travelled “rapidly” for relatively large
distances (Figure 5.36) down the face of the upstream slope whilst the bulk of the slide
mass was retained in the source area and travelled only limited distance (1.7 m and 4 m
respectively) at a likely much slower velocity. At Dam FD3 the toe of the slide mass
travelled up to 6 to 9 m distance down the 2H to 1V slope and onto the upstream bench.
At Forsythe dam the toe of the slide mass possibly travelled as much as 22 m down the
1.5H to 1V upstream slope to the upstream toe of the embankment.
The slope angle of the upstream slope is significant for the failures during
drawdown. For Type 1 slope failure geometries, break up of the toe region and “rapid”
travel of this broken up mass down the face of the upstream slope is likely for upstream
slope angles greater than or equal to 2H to 1V (or about 27 degrees). For the Type 1
and Type 2 slope failure geometries, analysis indicates that a correlation exists between
travel distance at the head of the slide and upstream slope angle for slides in medium
plasticity clays (Figure 5.40). However, there is not a lot of case data with which to
support the trend.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.79

40

Upstream Slope Angle (degrees) .


high plasticity clays and silts
medium plasticity clays
35
low plasticity clays and silts Medium plasticity
sandy and gravelly soils soils
30
Dam FD2 Dam FD3 Forsythe Sampna

25

20

15
High plasticity
soils
10
0.1 1 10 100
Total Deformation at Crest of Back-scarp (m)

Figure 5.40: Upstream slope angle versus total post failure deformation for slides in
embankment dams during drawdown.

5.6.4.3 Other Factors Affecting the Post Failure Deformation Behaviour of


Drawdown Triggered Slides in Embankment Dams
For the failures at Mount Pisgah, Dam FD3 and Old Eildon dam, factors other than or in
addition to material strain weakening on shearing and slope failure geometry were
considered significant in the post failure deformation behaviour. At Dam FD3 and
Mount Pisgah dam the presence of a concrete facing on the steep upstream slope (2H to
1V or steeper) was considered to contribute to the brittleness of the slide.
At Old Eildon dam, a concrete core-wall rockfill embankment with puddle clay zone
on the upstream side of the core-wall, the loss of support from the water load on the
upstream slope was considered to have contributed to the post failure deformation
behaviour. The initial failure was thought to have occurred during the drawdown of
1927 (prior to completion of construction). Additional rockfill was placed on the
upstream slope in 1928 whilst the reservoir was at close to full supply level, and during
this period it was noted that continued placement of rockfill was required in the slide
area due to ongoing subsidence (Knight 1938). On drawdown in 1929 a crest settlement
of 7 m occurred, followed by subsequent deformations during the later remedial works.
The post failure deformation on drawdown in 1929 was considered to have been a
reactivation due to the reduction of the water load on the upstream face and the change
in slope conditions from additional rockfill placement. Prior to the slide in 1929
deformations at the crest from previous slide movement were estimated at about 9 m.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.80

5.6.5 POST FAILURE DEFORMATION BEHAVIOUR OF FAILURES IN THE DOWNSTREAM


SHOULDER AFTER CONSTRUCTION

Of the thirteen case studies of slides in the downstream shoulder of embankment dams
that occurred post construction, the dominant embankment types (Table 5.7) consisted
of earthfill embankments (8 possibly 10 of 13 cases), in particular homogeneous
earthfill embankments (5 possibly 7 cases). For the zoned earthfill and zoned earth and
rockfill cases (1 case of each) the surface of rupture passed though the central core and
foundation. Table 5.16 provides a summary of the timing of the failure, embankment
type and geometry, material properties, drawdown causing failure and comments on the
slide itself. Further details are given in the tables in Section 1 of Appendix D.
Information on the materials types and properties, location of the surface of rupture
and post failure deformation records for these case studies is limited.
The post failure deformation and estimated peak velocity for the case studies is
presented in Figure 5.41. Of the deformation records, only in one case (Siburua dam)
are there records of deformation with time. For a further nine cases estimates of the
total deformations were available mainly from sections provided in the published
literature. Maximum post failure velocity ranges have been estimated for a number of
the cases based on the amount of deformation and reference to times over which the
deformation occurred. In general these estimates are very broad range and limited to
the IUGS (1995) velocity categories.
As for the slides in embankment dams during construction and during drawdown,
progressive strain weakening on shearing has the potential to result in post failure
acceleration and relatively large travel distance of the slide mass. This is particularly
important for slides in the downstream shoulder post construction due to the potential
for breach and uncontrolled release of the water storage of slides with large and
relatively rapid post failure deformation.
Assessment of factors that are significant in the deformation behaviour for slides in the
downstream shoulder post construction is somewhat subjective because of the limited
information on material properties and post failure deformation behaviour. Factors that
were found to be significant in the post failure deformation behaviour of the case studies
are:
• The potential for strain weakening due to particle orientation for near normally
consolidated high plasticity clays (Seven Sisters and Santa Ana Acaxochitlan dam).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.81

• Progressive failure. Progressive failure within the fissured clay foundation was
possibly a significant factor in the relatively high degree of brittleness of the failure
at Aran dam (as indicated by the large travel distance and high post failure velocity,
refer Figure 5.41).
• Slope failure geometry. The slope failure geometry and orientation of the surface of
rupture will affect the potential energy of the slide mass.
• Seepage into the deforming and dilating slide mass resulting in softening and
strength reduction (see below). Apart from Yuba dam, greater post failure
deformations and higher peak velocities were evident for the case studies where
seepage was considered significant in triggering the slide (Barton, Santa Ana,
Woodrat Knob, Fruitgrowers, Siburua, Great Western and Aran dams).
• Earthquake triggered liquefaction condition. For the slide at Yuba dam, triggered by
earthquake, liquefaction in either the foundation or poorly compacted earthfill may
have been a significant factor in the large and “rapid” post failure deformation
behaviour observed.

For the case studies where instability was considered to be directly related to
seepage and high pore water pressures (refer Section 5.5.3), the amount of post-failure
deformation was generally greater, from as small as 1.5 to 2 m up to 9 to 11 m. The
maximum post-failure velocity was estimated to range from slow (Great Western) to
rapid, with most cases estimated to be in the moderate category. The failures at Santa
Ana, Aran and Fruitgrowers dams were estimated to be in the moderate to rapid and
rapid category. Dilation on shearing and softening due to infiltration of free water into
the slide mass may have been significant for these failures.
For the case studies where, apart from Yuba dam, instability was considered not
directly related to seepage and high pore water pressures (Harrogate, Seven Sisters and
possibly Park Reservoir dams), the post failure deformation of the slide mass was less
than 2 m and peak post-failure velocities were estimated to be in the slow or slow to low
end of moderate categories.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.82

Table 5.16: Summary of the case studies of slides in the downstream shoulder of embankment dams that occurred post construction.
Material Properties (in zone of
Years Slope *1 Embankment Details
Name of Year of surface of rupture) Location of
since Failure Comments
Dam Failure Type Hgt Slope Embankment *2 Foundation *2 Back-scarp
Const d Geometry
(m) (o)
Park 1969 60 Type 1 E’fill – 24.4 33.7 SC – clayey sand. No (failure through Crest, centre Rotational slide in the upper part of the
Reservoir rock toe formal compaction (?). embankment only) steep downstream slope (above the rockfill
(2,0,0) toe).
Arroyito 1984 6 Type 2 E’fill - 20 21.8 GP/GW – well (failure through Downstream Shallow rotational slide, possibly a
filters compacted sandy embankment only) slope sloughing failure. Seepage out of the slope.
(1,0,2) gravels, < 10% fines
Barton 1922 12 Type 1/2 (?) Homog 12.2 33.7 CH – silty clays. No (failure through Crest, Rotational slide in the steep downstream
(0,0,0) formal compaction. embankment only) downstream slope. Seepage on the downstream face.
edge
Harrogate 1951 81 Type 2 Puddle 8.8 27.8 CL – Boulder clay of Deep clay profile, Crest, Slow rotational slide. May have extended
E’fill medium plasticity. No stiff to very stiff. downstream into foundation. Triggered by rainfall.
(8,0,0) formal compaction. Surface soil not edge Deep shrinkage cracks in the embankment.
removed.
Woodrat 1961 5 unk Homog 26 29.7 unk (suspect clayey and unk (possibly Downstream No details. 9 m of post failure deformation
Knob (0,0,0 ?) low permeability) through fndn) slope at moderate rate. Steep downstream slope.
Fruit- 1937 39 (2 since Type 1/2 (?) Homog 11 26.6 CL – silty clay (residual SC - clayey sands, Crest, centre Likely rapid rotational slide. Occurred
growers raise) (0,0,0) shales) of medium alluvial (possibly when reservoir was at its highest historical
plasticity. No formal through fndn) level. Seepage observed prior to the
compaction. downstream raise in 1935.
Santa Ana 1952 21 Type 2 Homog 12 26.6 CH – silty clays. CH/OH – organic Upstream slope Deep seated compound (?) slide in the
(0,0,0) clay layer over soft downstream slope that extended back to the
clays. No fndn clean upstream slope. Moisture observed on
up. downstream slope in month prior to failure.
Siburua 1964 7 Type 5 E’fill - 14.4 26.6 CL – shaly clays of CL – shaly clays of Downstream Shallow compound slide through the
filters medium plasticity. medium plasticity. slope embankment and foundation in the lower
(1,0,1) Rolled, well compacted Possible favourable downstream slope. Slide initiated when the
relict bedding. reservoir was at full supply level.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.83

Table 5.16: Summary of the case studies of slides in the downstream shoulder of embankment dams that occurred post construction (Sheet 2).
Material Properties (in zone of
Years Slope *1 Embankment Details
Name of Year of surface of rupture) Location of
since Failure Comments
Dam Failure Type Hgt Slope Embankment *2 Foundation *2 Back-scarp
Const d Geometry
(m) (o)
Seven 1949 to 0.3 to 13 Type 5 Zoned 7.3 21.8 CH – clay, placed well CH – clay, firm to Crest, upstream Deep seated (for low embankment height)
Sisters 1956 E&R wet of OMC. Rolled. stiff. Fissured to 2.1 edge rotational slide through the soft clay
m. foundation.
Great 1958 51 (1 since Type 2 Homog 18.6 unk Rolled earthfill Alluvial clays Upstream slope Deep seated compound (?) slide that
Western raise) (0,0,0 ?) (≈ 26 ?) extended back to the upstream embankment
slope. Close to highest reservoir level in
embankment history at the time of failure.
Aran 1978 1 Type 2/5 (?) Zoned 30.3 23 CH – clay, rolled CH – alluvial Crest, Deep seated compound (?) slide through the
E’fill fissured clays. downstream fissured clay foundation. Rapid post failure
edge deformation.
Lake 1943 60 Type 2 Homog 16.2 26.6 CL/SC – low plasticity CL/SC – as per Downstream Shallow rotational (?) slide that passed
Yosemite (0,0,0) gravelly sandy clays embankment slope through upper foundation. Possibly
and clayey sands. No material. Referred to triggered by rainfall.
formal compaction. as cemented.
Yuba 1951 41 (2 since Type 2 Homog 7.6 29.7 CL – sandy clays of Alluvial clays Crest, centre Compound (?) slide triggered by
raise) (0,0,0) medium plasticity. No earthquake. Seepage on the downstream
formal compaction. slope. Likely rapid post failure
deformation.
Notes: Hgt = height Const d = construction unk = unknown OMC = Standard optimum moisture content
Homog = homogeneous earthfill embankment E’fill – filters = earthfill embankment with filters
E’fill – rock toe = earthfill embankment with rock toe Zoned E’fill = zoned earthfill embankment
Puddle E’fill = puddle core earthfill embankment Zoned E&R = zoned earth and rockfill embankment
*1 Slope failure geometries defined in Figure 1.3 in Section 1.3 of Chapter 1.
*2 Soil classification to the Australian soil classification system (AS 1726-1993)
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.84

(a) 14
Clay (CH) Clays (CL) Clays (CL) Clayey
high plasticity medium plasticity low plasticity Sand
12
Total Post Failure Deformation (m)

10

at head of back-scarp
8
at toe of slide

?
2

Lake Yosemite
Santa Ana

Seven Sisters

Aran

Siburua
Harrogate

Yuba

Great Western
Barton

Fruitgrowers

Woodrat Knob
(b) 1.E+08
Clays (CH) Clays (CL) Clays (CL) Clayey Very
high plasticity medium plasticity low plasticity Sands
Rapid
1.E+07
Maximum Velocity (mm/day) .

? ?
1.E+06
Rapid

1.E+05
? ?
1.E+04
Moderate
?
1.E+03
? ? ?
1.E+02 Slow
Estimate from monitoring data Broad range estimate
1.E+01
Lake Yosemite

Park Reservoir
Santa Ana

Seven Sisters

Aran

Siburua
Harrogate

Yuba

Great Western
Barton

Fruitgrowers

Woodrat Knob

Figure 5.41: Post construction slides in the downstream shoulder; post failure (a)
deformation, and (b) estimated peak velocity.

5.6.5.1 Potential for Dam Breach due to Instability of the Downstream Slope
An important aspect of post construction slides in the downstream shoulder is the
potential for large deformation on sliding that could potentially result in a breach and
uncontrolled release of water from the reservoir. The slide at Santa Ana dam resulted in
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.85

breach of the embankment, at Fruitgrowers dam it was a significant factor in the


eventual breach (refer Section 5.2.1) and at Great Western dam a breach was only
narrowly averted. Importantly, the reservoir was at or close to full supply level at the
time of the failure in these three cases and a number of the other cases where instability
was triggered by seepage and high pore water pressures.
The case studies of failures during construction (Section 5.6.3) indicates that there is
a greater likelihood for extension of the back-scarp to incorporate the opposite slope
where the surface of rupture involves the foundation. In eight of the ten cases involving
the foundation the back-scarp extended back to between the central region of the crest
and the opposite slope. This compares to three of seven cases involving the
embankment only (excluding the cases of lateral bulging). The strength of the
foundation in comparison to the strength of the embankment material/s is significant in
the likelihood of involvement of the foundation in the failure.
For the post construction slides in the downstream shoulder the case study evidence
is not as clear as for the failures during construction. In three of the six cases where the
slide back-scarp extended to or upstream of the embankment centreline (refer Table
5.16), the surface of rupture passed through the foundation (Santa Ana, Great Western
and Seven Sisters dams). In two cases it was unsure if the foundation was involved in
the failure (Fruitgrowers and Yuba dam) and at Park Reservoir dam the surface of
rupture was through the embankment only (the small slide was confined to the upper
part of the steep down-slope). Of the three slides known to incorporate the foundation,
in two cases (Santa Ana and Seven Sisters dams) the foundation was of high plasticity
clays, and in the case of Santa Ana dam the back-scarp extended back to a level below
that of the reservoir and a breach subsequently occurred. At Great Western dam the
foundation was of alluvial clays of unknown plasticity, and although the back-scarp
extended back to a level below that of the reservoir on the upstream slope of the crest, a
breach was averted by drawing down the reservoir and stabilising the downstream toe of
the embankment.
Therefore, embankments founded on weak or high plasticity clay foundations are
potentially susceptible to deep-seated failures that could potentially lead to an
uncontrolled breach from the reservoir. Higher post failure velocities are likely with
increased brittleness of either the materials and/or the slide mechanics, and where the
trigger for failure is seepage related.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.86

5.7 P OST F AILURE D EFORMATION BEHAVIOUR - CUT SLOPES IN


HEAVILY OVER-CONSOLIDATED CLAYS
This section presents the post failure deformation behaviour for the case studies of
failures in cut slopes in heavily over-consolidated high plasticity clays. Data on travel
distance and travel velocity of the case studies is presented, and the factors affecting the
deformation behaviour of the slide mass are discussed, with greater detail given for
several case studies to highlight some of the factors affecting deformation behaviour.

5.7.1 SUMMARY OF FACTORS AFFECTING THE POST FAILURE DEFORMATION

Progressive failure is a significant factor in the mechanics of failure of cuts in high


plasticity heavily over-consolidated clays as discussed in Section 5.4.1. It is also
significant in the mechanics of post failure deformation. The brittleness of the slide
post failure is dependent, amongst other factors, on the strength loss on shearing from
the shear strength at failure. Therefore, if at failure the strength of the soil mass is close
to peak strength (i.e. progressive failure was not greatly significant in the mechanics of
failure), such as undrained failure of a very steep cut, then the failed slide mass has a
high potential brittleness due to the large strength difference between at failure and
residual strength. Conversely, if the average shear strength along the surface of rupture
at failure is close to the steady state strength (i.e. progressive failure was significant in
the mechanics of failure) then the potential brittleness of the failed slide mass is much
lower and due to the smaller difference in strength between at failure and residual
strength.
Other factors contributing to the brittleness of the slide mass at failure and the post
failure deformation behaviour of cuts in heavily over-consolidated high plasticity clays
are:
• Internal brittleness in the slide mechanics.
• Brittleness due to three-dimensional effects and release of the slide on the lateral
margins, such as formation of shear surfaces through retaining walls.
• Orientation of the surface of rupture. For most Type 1 and Type 2 slope failure
geometries the surface of rupture was of compound shape with basal translational
component and rotational back-scarp, such as the failure in London clay at Northolt
in 1955 (Figure 5.42).
• Slope failure geometry.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.87

Figure 5.42: Slide in the cut slope in London clay at Northolt in 1955 (Skempton 1964)

5.7.2 SUMMARY OF THE POST FAILURE DEFORMATION AND VELOCITY OF THE


FAILURE CASE STUDIES

Table 5.17 and Table 5.18 present a summary of the slope geometry and post-failure
travel distance and velocity of the failure case studies within cut slopes in London clay
and Upper Lias clay respectively. Further details on the individual cases are given in
Section 2 of Appendix D.
The post failure deformation records for most of the case studies analysed is limited.
Monitored deformation records were only available for several cases. For most cases
the deformation was estimated from pre and post failure sections of the slides. Post
failure velocities were estimated from the total deformation and qualitative information
on the time over which the deformation occurred. These velocity estimates are broad
range and have generally been restricted to the IUGS (1995) velocity categories unless
sufficiently detailed information was available for more precise estimation.
Figure 5.43 and Figure 5.44 present plots of the post failure deformation
(deformation plotted as a percentage of the slide length measured as the straight line
distance from the head to the toe of the slide) and estimated peak post failure velocity
for the case studies of cuts in London clay. For the Type 1 and Type 2 slope failure
geometries the post failure deformation and velocity typically increased with increasing
cut slope angle. But, the figures also show that a significant change in deformation
behaviour (for both deformation and velocity) is evident for the Type 1 and Type 2
slope failure geometry. This trend is more clearly evident in Figure 5.45 and Figure
5.46 where the post failure deformation and velocity are plotted against cut slope angle.
The group of slides of rapid post failure velocity and greater deformation occur in the
steeper cut slopes when the slides are separated on slope failure geometry.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.88

Table 5.17: Cut slope failures in London clay.


Slope Post-failure
Age Geometry Deformation
Name SFG *1 Comment
(yrs) Hgt Slope Deformation Velocity*2
(m) (º) (m) Category
Hadley Wood 70 13.6 13.5 0.9 (c), 1.8 (t) Slow Possible reactivation of deformation.
Grange Hill 50 17.4 17.8 2 (c & t) Slow Occurred in winter, suspect slow.
West Acton A 44 4.7 18.4 0.36 (c), 1.5 (t) Slow Limited deformation, suspect slow
Northolt 19 10.1 21.8 1.2 (c), 0.8 (t) Slow to Compound slide in slope above
1
moderate retained section.
Grove Park 98 9.4 22.5 3.3 (c), 10.8 (t) Rapid At toe slide travelled 10.8 m on 22.5
degree slope.
New Cross 3 23 33.7 10 (c), 31 (t) Rapid Rapid compound failure in steep cut.
Crews Hill 47 5.3 16.2 1.2 (c), 1.5 (t) Slow Possibly a reactivated translational
slide
West Acton B 44 4.8 18.4 0.41 (c), 2.5 (t) Slow Limited deformation, suspect slow.
Whitstable 99 11.4 18.4 1.5 (c), 2 (t) Slow Deformation occurred over a period
of several weeks to months.
Sudbury Hill 46 7 18.4 2.2 (c), 1.6 (t) Slow Deformation in winter over several
years.
Fareham 59 9.5 19.5 1.5 (c), 4 (t) Slow Deformation over several years,
reactivated in 1967.
Althorne 60 8.8 20.5 2 (c), 3 (t) Slow Suspected reactivation.
Cuffley 38 12 20.7 2.8 (c), 3.5 (t) Slow Seasonal winter deformations over at
least 6 years.
2
Dedham 112 10.2 22 2.1 (c), 5 (t) Moderate Continued slow deformation when
toe of slide cut back.
Kingsbury S1 16 6.2 23.5 0.5 (c), 1.0 (t) Slow Limited deformation, suspect slow.
St. Helier 22 7.3 25.3 2.3 (c), 2.5 (t) Slow to Deformation over at least 4 years.
moderate
Isle of 6 12 38 12.5 (t) Rapid (?) Steep cut slope, suspect rapid
Sheppey A velocity of deformation.
Isle of 8 10 41 9 (t) Rapid (?) Steep cut slope, suspect rapid
Sheppey B velocity of deformation.
Uxbridge 37 9 43 5 (t) Rapid Failure of temporary vertical
1937 excavation, suspect rapid.
Bradwell 5 days 14.8 56 unk Rapid Steep cut, rapid deformation.
Upper 81 5.3 25 1.3 (c) to 0.7 Slow Initial deformation in 1951.
Holloway (t) Reactivation of slide (new retaining
wall) in 1953-55.
Wembley Hill 13 17 26 6 (c & t) Rapid 6 m deformation over 0.5 hour
5
period, only slight deformation
(retained
thereafter.
cuts)
Wood Green 55 11.3 28 0.9 (t) Slow Limited slow deformation.
Kensal Green 29 6.1 29.7 0.3 (t) Slow Limited slow deformation.
Uxbridge 54 10.2 32 0.7 (t) Slow Deep-seated failure.
1954
Age = age of cut slope at failure Hgt = cut slope height Slope = cut slope angle
(c) = crest deformation (t) = deformation at toe unk = unknown
*1 SFG = Slope failure geometry. Refer Section 1.3 for definitions.
*2 IUGS (1995) velocity category (refer Table 1.1 in Section 1.3).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.89

Table 5.18: Cut slope failures in Upper Lias clay.


Slope Post-failure Deformation
Age Geometry SFG
Name 1 Comment
(yrs) Hgt Slope * Deformation Velocity*2
(m) (º) (m) Category
Ardley 52 15 22 1 0.5 (c), 0.8 (t) - Slide in slope above retaining
wall following heavy storms.
Wothorpe B 5 6.1 18.5 3.5 (c), 2 (t) Deformation initiated in winter of
slow 1964/65 and continued over
several years.
Seaton 67 9.3 20.8 2.0 (c), 2.2 (t) - Translational slide in brecciated
clays.
Stowehill B 122 8.2 23 0.6 (c), 1.4 (t) - Rotational slide that partly
encompassed earlier slide (56
years earlier).
Heyford 125 9.1 24 2.0 (c), 2.3 (t) - Slide in winter following very
heavy December 1960 rainfall.
Hunsbury Hill 77 16.4 24.5 7.6 (c), 1 (t) - Reactivated slide deformations
Tunnel (1954) over many years. Large crest
deformation as a result of
2 continued trimming back of toe.
Stowehill A 66 10.2 24.8 2 (c), 1 (t) - Initial Stowehill slide in 1901.
Culworth A 61 8 26.5 2 (c & t) - Slide occurred in the winter of
1957/58 following a very wet
summer.
Culworth B 61 8 26.5 2 (c), 1.6 (t) - Slide occurred in the winter of
1957/58 following a very wet
summer.
Charwelton 40 9.6 27.2 1.4 (c), 1.3 (t) slow Slide active for more than 1
week, therefore slow.
Barrowden 83 4.3 28.8 0.55 (c), 0.5 (t) - Deep-seated rotational slide.
Wellingborough 105 10 31 unk - Limited information on slide.
Station

Age = age of cut slope at failure Hgt = cut slope height Slope = cut slope angle
(c) = crest deformation (t) = deformation at toe unk = unknown
*1 SFG = Slope failure geometry. Refer Section 1.3 for definitions.
*2 IUGS (1995) velocity category (refer Table 1.1 in Section 1.3).

No such correlation is evident for the slides in cut slopes of London clay of Type 5
slope failure geometry, indicating that cut slope angle is not greatly significant for these
failures.
For the cut slopes in Upper Lias clays, estimates of the post failure velocity were
possible for only two cases. Both were estimated to be in the slow IUGS category. The
limited post failure deformation of all of the cut slope failures (Table 5.18) suggests it is
likely that the post failure velocity is within the slow to lower end of the moderate range
for them all. In addition, no travel distance correlation is evident with respect to cut
slope angle. This is likely to be partly due to the limited range of cut slope angles for
the case studies, which ranged from 18 to 31 degrees.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.90

Post-failure Deformation (% of slide length) .


100%
Type 1 slope Type 2 slope Type 5 slope
failure geometry failure geometry failure geometry

increasing cut increasing cut increasing cut


slope angle slope angle slope angle

10%

Deformation at head of slide


Deformation at toe of slide
1%

Upper Holloway
Grange Hill

Dedham
Northolt

Sudbury Hill
Crews Hill

Fareham

Uxbridge - 1937

Bradwell

Kensal Green

Uxbribge - 1954
Hadley Wood

Wembley Hill

Wood Green
Cuffley
Grove Park

New Cross

Althorne
Whitstable

St. Helier
Kingsbury - S1

Isle of Sheppey - A

Isle of Sheppey - B
West Acton - A

West Acton - B

Figure 5.43: Post-failure slide deformation for cut slopes in London clays.

1.E+07
Estimated Peak Velocity (mm/day) .

Type 1 slope Type 2 slope Type 5 slope


failure geometry failure geometry failure geometry
1.E+06
increasing cut
Rapid
slope angle
1.E+05

increasing cut
1.E+04 slope angle
increasing cut
Moderate
slope angle
1.E+03

1.E+02 Slow

1.E+01
Upper Holloway
Grange Hill

Dedham
Northolt

Sudbury Hill
Crews Hill

Fareham

Uxbridge - 1937
Bradwell

Kensal Green

Uxbribge - 1954
Hadley Wood

Wembley Hill

Wood Green
Cuffley
Grove Park

New Cross

Althorne
Whitstable

St. Helier
Kingsbury - S1

Isle of Sheppey - A

Isle of Sheppey - B
West Acton - A

West Acton - B

Figure 5.44: Estimated peak post failure velocity of slides in cut slopes in London clays.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.91

60
Slide deformation restricted by Type 1 slope failure geometry - London clay
Cut Slope Angle (degrees) . the width of the excavation Type 2 slope failure geometry - London clay
50 Type 2 slope failure geometry - Upper Lias clay

40

rapid velocity
30 slides

20

10 "slow " velocity


slides

0
0% 10% 20% 30% 40% 50% 60% 70% 80% 90%
Deformation at Toe of Slide (% of slide length).

Figure 5.45: Cut slope angle versus post-failure deformation at the slide toe for cut
slopes in heavily over-consolidated clays of Type 1 and Type 2 slope failure geometry.

60
Type 1 slope failure geometry - London clay
Type 2 slope failure geometry - London clay
50
Cut Slope Angle (degrees) .

Type 5 slope failure geometry - London clay


Type 2 slope failure geometry - Upper Lias clay
40

30

20

10
slow moderate rapid

0
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06
Estimated Post-Failure Velocity of Slide (mm/day)

Figure 5.46: Cut slope angle versus estimated post-failure slide velocity for cut slopes in
heavily over-consolidated clays.

The deformation behaviour within the three types of slope failure geometry (Type 1,
Type 2 and Type 5) is discussed in the following sections. Factors influencing the post
failure deformation behaviour are discussed in greater detail under the section on Type
2 slope failure geometry, the most dominant category for the failure case studies in cut
slopes.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.92

5.7.3 FAILURES IN CUT SLOPES – TYPE 2 SLOPE FAILURE GEOMETRY

For the Type 2 slope failure geometries the post failure deformation behaviour, from
analysis of the case studies, indicates:
• At cut slopes less than about 28 to 35 degrees the slide mass remains virtually intact
and within the source area on sliding. Post failure deformation of the slide mass is
generally limited to less than 2 to 4 m, and the peak velocity is likely to be in the
slow to lower end of the moderate range. Post failure deformations generally occur
over a period of days to weeks and are readily reactivated during adverse weather
conditions.
• At cut slopes greater than about 30 to 35 degrees, the post failure travel distance of
the toe region of the slide is generally much greater (50 to 70% of the slide length,
Figure 5.45) and post failure velocities typically in the rapid range (Figure 5.46).
The post-failure deformation generally occurs over a short time period (minutes to
hours in most cases depending on the steepness of the cut slope). Continued
deformations after the initial event are likely to be very limited.

For the Type 2 cut slope failures in London clays in slopes less than 28 to 30
degrees (i.e. the “slow” slides), the age of the slope at failure ranged from 16 to more
than 100 years after excavation. For these failures progressive failure is likely to have
been a significant factor in the mechanics of failure.
For these “slow” slides of Type 2 slope failure geometry (and also Type 1) the post-
failure deformation at initial (or first time) failure occurred “slowly” over many days.
Deformations of the slide mass were reactivated during subsequent wet winters or
seasonally wet periods in the years following the initial failure (James 1970), and
generally at lower velocities than at the initial failure. This seasonal reactivation
indicates that the failed slope is of marginal stability and triggered by seasonally high
pore water pressure conditions in the slope.
Two-dimensional limit equilibrium analysis of some of these “slow” slides (Table
5.19 in Section 5.8.1) indicated that the residual factor of safety was relatively low (0.55
to 0.8). This would suggest that the initial slide has a potentially high degree of
brittleness post failure. However, the mechanics of post failure deformation as observed
would suggest that the kinetic energy of the slide mass at any time is relatively small
(i.e. most of the available potential energy is consumed as frictional energy or by other
means).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.93

The factors restricting these “slow” slides from reaching higher velocities at initial
failure (and therefore greater travel distances) are thought to include:
• Restraint on the lateral margins in the form of resistance to development of a shear
surface, including three-dimensional effects such as changes in slope geometry.
• Strain hardening on shearing in these over-consolidated clays and development of a
reduction in pore water pressure due to dilation. The effect of strain hardening on
the deformation behaviour would be to retard the deformation of the slide mass.
• Change in geometrical shape with deformation of the slide mass, i.e. the slide
deforms to a more stable configuration for a given set of strength parameters.
• Visco-plastic properties of these high plasticity soils.

All explanations are considered to have some merit in the overall mechanics of post
failure deformation for these “slow” slides. The strain hardening response is considered
to have greater merit in terms of being able to explain the subsequent reactivation of
deformation under climatic conditions less adverse than what triggered the initial slide.
Reduced pore water pressures are developed in the region of the surface of rupture due
to the propensity of the soil to dilate on shearing. Moisture will then migrate to these
regions of low pore water pressure due to pore water pressure variation set up in the
slope post the failure, and with time, result in softening and dilation in the region of the
surface of rupture. The consequent reduction in the shear strength properties means that
the slope is of limiting equilibrium during subsequent seasonally wet periods and
deformation is reactivation. Equilibration of pore water pressure variations in the slope
are likely to occur slowly in these high plasticity clays due to their low permeability.
The cut slope failures in slopes greater than 30 to 35 degrees in London clays
generally occurred in a matter of days to years after excavation (upper limit of 8 years
for the case studies). The rapid post failure velocity and greater deformation of the slide
mass for these slides is indicative of a greater conversion of available potential energy
to kinetic energy than for the “slow” slides. These “rapid” slides tended to break up to a
greater degree post failure indicating a greater amount of energy was used in
disaggregation of the slide mass.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.94

5.7.4 FAILURES IN CUT SLOPES – TYPE 1 SLOPE FAILURE GEOMETRY

For the Type 1 slope failure geometries the post failure deformation behaviour, from
analysis of the case studies, indicates:
• At cut slopes less than about 22 degrees, the slide mass remains virtually intact on
sliding within the slide source area. Limited post failure deformation (less than
about 1 to 2 m) at a velocity in the slow to lower end of the moderate range is
generally observed. Deformations generally occur over a period of days to weeks
and are readily reactivated during adverse weather conditions.
• At cut slopes greater than about 22 degrees, the toe region of the slide will
potentially break up and travel at a moderate to rapid velocity and travel for
relatively large distances (over a short time period) on the lower portion of the cut
slope. The portion of the slide that does not exit the failure bowl remains virtually
intact and, as observed, the travel distance is much lower than at the toe of the slide.
Further deformations after the initial event are likely given the potential for
additional material to move out of the failure bowl and onto the cut slope below.
• With increasing cut slope angle above about 22 degrees, the volume of material
likely to evacuate the source area and travel “rapidly” down the cut slope below is
likely to increase. At Grove Park (cut slope of 22.5 degrees) only a small volume of
the slide mass evacuated the source area compared with the much larger volume at
New Cross (Figure 5.47) where the cut slope angle was much steeper at 34 degrees.

Figure 5.47: Failure in the cut slope in London clay at New Cross in 1841 (Skempton
1977)

The brittleness of the slide mass at failure is considered a significant factor in the
relative volume of material that evacuates the source area. For the failure at New Cross,
which occurred three years after excavation, the average available shear capacity along
the surface of rupture at failure and therefore potential for post failure strain weakening
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.95

on shearing was likely to have been significantly greater than at Grove Park where
failure occurred 98 years after initial excavation.

5.7.5 FAILURES IN CUT SLOPES – TYPE 5 SLOPE FAILURE GEOMETRY

For the failure case studies of Type 5 slope failure geometry (all retained cut slopes in
London clays), the cut slope angle is not a significant factor in the post failure
deformation behaviour (Figure 5.46). In four of the five case studies the post-failure
deformation was less than 1.5 m and post failure velocity was estimated to be in the
slow category (Table 5.17). In contrast, the slide at Wembley Hill (Figure 5.48)
travelled a distance of 6 m, most of this within a 30 minute period, and therefore at a
rapid velocity.

Figure 5.48: Failure in the retained cut slope in London clay at Wembley Hill
(Skempton 1977)

The factors controlling the mechanics of post failure deformation behaviour are as
previously discussed for the Type 2 slope failure geometries (Section 5.7.3). Conditions
resulting in increasing kinetic energy of the slide mass will result in acceleration to
higher velocities and greater travel distances.
For the slide at Wembley Hill factors influencing the post failure deformation
behaviour potentially include:
• Greater inertia of the slide mass. The volume and cross sectional area of the
Wembley Hill slide (volume estimated at 70000 m3) were significantly greater than
for the other four Type 5 slides.
• Brittleness associated with release on the lateral margins. At Wembley Hill loud
reports were heard at the time of failure, most likely associated with failure through
the retaining wall on the margins. Of the other four Type 5 case studies, cracking in
the retaining wall was reported at Uxbridge and it is suspected that similar limited
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.96

damage to the retaining wall was likely at the other retained cut slope sites given the
limited deformation observed.
• Timing of the initial recognition of the slide to the stage of failure and whether or
not any remedial works were undertaken to ameliorate the stability of the slope.

Deformation behaviour at Uxbridge in 1954


The slide at Uxbridge in 1954 (Figure 5.49) is considered to be typical of the “slow”
post failure deformation behaviour of slides of Type 5 slope failure geometry in retained
cut slopes. The post failure deformation of the slide developed slowly over many days
before accelerating as shown in Figure 5.50. The first signs of instability (Watson
1956) were observed in late September of 1954, some 60 days prior to the monitoring
shown, in the form of cracking in the pavement above the cut slope. By the 10th of
November 1954 further deformation was observed in the head region of the slide as well
as deformation in the toe region (a crack in the retaining wall was observed to have been
getting worse and the track drain beyond the retaining wall was not functioning
properly). By the 22nd of November the railway officials were unable to maintain the
super-elevation of the track in the vicinity of the slide indicating heave deformations in
this area. Monitoring points were then established on the retaining wall in the toe
region of the slide. The records show an increase in the post failure velocity of the slide
over a period of several days (from < 10 mm/day to about 150 mm/day).
Restraint on the lateral margins (mainly from the retaining wall, but also from the
changing slope topography) and passive resistance at the toe of the retaining wall are
thought to have been influential in the “slow” post failure velocity of the slide. Given
the length time over which the slide developed (a six to eight week period), it was
possible for remedial measures to be undertaken to slow down and stop the deformation
of the slide as was observed (Figure 5.50). These measures included excavation works
at the head of the slide, filling works at the toe of the retaining wall and strutting of the
retaining wall (Watson 1956).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.97

Figure 5.49: Failure in the retained cut slope in London clay at Uxbridge in 1954
(adapted from Henkel 1956)

500 200
Strutting of retaining wall Point 1 Point 2
started on 27/11/54
Displacement of Retianing Wall (mm)

400 (Day 4) 160


Displacement

Velocity (mm/day) .
300 120

200 80

Velocity
100 40

0 0
0 2 4 6 8 10 12 14
Time (Days from 23 November 1954)
Figure 5.50: Monitored post-failure deformation of the slide in the retained cut in
London clay at Uxbridge in 1954 (adapted from Watson 1956).

5.8 P REDICTION OF P OST-F AILURE TRAVEL DISTANCE

5.8.1 KHALILI ET AL (1996) MODEL FOR INTACT SLIDES

The Khalili et al (1996) model for estimating the post-failure deformation of intact
slides has been used to analyse the post failure deformation behaviour of a number of
the failure case studies in embankment dams and cut slopes in heavily over-consolidated
high plasticity clays. The model is described in Section 4.3.5.2 of Chapter 4. In
summary, the model considers two modes of failure; a “rapid” model based on the
principle of conservation of energy (i.e. a simplified lumped mass model), and a “slow”
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.98

model based on equations of equilibrium between the driving and resisting forces of the
slide.
The failure model is shown in Figure 5.51. ?i and ?f are the initial and final
positions of the centre of gravity of the slide mass. Note that the terms “rapid” model
and “slow” model defined and used by Khalili et al (1996) are different to the
definitions of “rapid” slides and “slow” slides used here (refer Section 1.3 for
definitions) and the definitions of rapid and slow post failure velocity defined by IUGS
(1995).

Figure 5.51: Failure model for post-failure deformation (Khalili et al 1996)

Analysis of the post-failure deformation using the Khalili et al (1996) method has been
undertaken on twenty-two case studies. In most cases the slide mass remained virtually
intact during sliding. A summary of the cases analysed is given in Table 5.19 and
included:
• Four cut slopes in heavily over-consolidated, high plasticity Upper Lias clays (all of
Type 2 slope failure geometry). The slope height of the four cases ranged from 4 to
10 m and the cut slope angle from 18.4 to 28.8 degrees.
• Nine cut slopes in heavily over-consolidated, highly plasticity London clays. The
cut slope height ranged from 5 to 17 m and cut slope angle from 18.4 to 38 degrees.
Seven of the cases analysed were of Type 2 slope failure geometry and one each of
Type 1 (Northolt) and Type 5 (Wembley Hill).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.99

• Five cases of slope instability within embankment dams. These cases included one
failure during construction (Carsington dam), three failures on drawdown (Dam
FD2, San Luis and Old Eildon dams) and one post construction slide in the
downstream slope (Siburua dam).
• Three cases of slope instability of embankments on soft clay foundations. In two
cases (St Alban A and Portsmouth) the foundation comprised low plasticity
sensitive clays and in the remaining case (Rio) the foundation comprised high
plasticity clays of low sensitivity. Details on the analysis of these three case studies
are given in Section 4.3.5.2 (of Chapter 4).
• One case of slope instability in a part fill part quasi-natural road embankment (Case
RF1) of 15 m height and average slope angle of 25 degrees. The initial failure
occurred shortly after completion of road widening and was triggered by a heavy
and prolonged period of rainfall. Subsequent reactivation occurred during wet
periods over the following 4 years. The compound slide had a basal translational
surface of rupture along a pre-sheared surface in the foundation, possibly associated
with part of the toe region of an ancient slide, and rotational back-scarp through
compacted filling.

Further details of the individual embankment dam and cut slope case studies
analysed are given in Sections 5.4 to 5.7 and in the tables in the Appendix D. Case
studies of Type 1 slope failure geometry where break up and “rapid” travel of the toe of
the slide mass occurred were not analysed using the Khalili et al (1996) model because
is not appropriate for slides where separation of the slide mass occurs.
The methodology used for analysis of the case studies was as follows:
(i) Limit equilibrium analysis (using the computer program SLOPE-W) to estimate
the centre of rotation of the slide and therefore the radius, r, and initial position of
the centre of gravity, ?i. This was done by approximating the circular slide
surface to the known surface of rupture of the actual slide. Materials parameters
and pore water pressure conditions used for this part of the analysis were those
giving close to a factor of safety of unity.
(ii) Limit equilibrium analysis to determine the factor of safety at residual strength
(FSresidual). This was carried out using the known surface of rupture and
Morgenstern Price force equilibrium method for cases with a non-circular surface
of rupture.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.100

(iii) Estimation of ?f from the post failure deformation of the slide mass and
approximated circular surface of rupture (determined in step i), and calculation of
θ i − θ f and 1 − θ f θ i .

In most cases the actual surface of rupture was classified as compound (i.e. non-
circular) as opposed to rotational upon which the methodology is based. In a number of
non-circular cases the fitted circular analysis was compared to a non-fitted circular
analyses. The results indicated that, in most cases, there was no significant effect on the
initial angle of the centre of gravity (?i); however, the fitted analysis did tend to result in
a significantly greater radius for the slip circle (r in Figure 5.51).
Residual strength parameters were assessed from available laboratory test results.
For the embankment dam and fill on soft ground cases this was usually from site
specific laboratory tests. For the cut slopes in London and Upper Lias clays it was from
published representative parameters of residual strength (Skempton et al 1969; Chandler
1976; Skempton 1977; Chandler 1984b). Drained residual strength parameters were
used for heavily over-consolidated soils such as rolled, well-compacted earthfills and
the over-consolidated natural soils. For normally and near normally consolidated soils
residual undrained strength parameters were used. These included the wet placed core
materials of Carsington and Old Eildon dams, and the soft foundations of the fill on soft
ground case studies (Rio, St. Alban A and Portsmouth).
The use of undrained strength parameters for the normally and near normally
consolidated soils obviated the need for assessment of pore water pressures. However,
for the soil types analysed in terms of drained strength it was necessary to determine
pore water pressures. For these cases the results of the analysis were sensitive to the
pore water pressure parameters/profiles used, and there was the added uncertainty
associated with shear induced pore water pressures. Where possible, piezometric
records were used for assessment of the pore water pressure conditions and these were
generally available for most of the case studies analysed. In cases where no piezometric
records were available (Wembley Hill, Dedham, Althorne, Kingsbury S1, St. Helier,
West Acton B and Isle of Sheppey A) a pore water pressure coefficient, ru, was
estimated based on the age of the cut slope from correlations given by Skempton (1977)
(Figure 5.10).
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.101

The results of the deformation analysis are presented in Figure 5.52 and summarised
in Table 5.19. For two cases, 21 (Portsmouth) and 17 (Dam FD2), the actual location of
the plotted point in Figure 5.52 was thought to differ from that calculated. In the case of
Portsmouth the residual factor of safety was estimated based on residual undrained
shear strengths determined by field shear vane. The remoulded vane strength was
considered to significantly under-estimate the large strain undrained shear strength and
therefore result in low predictions of the residual factor of safety (refer Section 4.3.5.2).
The actual location of Portsmouth on Figure 5.52 is considered to be located toward the
mid range between the “slow” and “rapid” model estimates as shown.
In the case of Dam FD2 the post failure deformation records used in the analysis
were only those recorded during the 1994 drawdown, which amounted to 430 mm. The
deformation that occurred during the initial slide event, thought to be during the 1993
drawdown, was not included as no indication of the amount of deformation was given.
The actual location of point No. 17 is therefore likely to be further to the right toward
the “slow” model as indicated by the arrow in Figure 5.52.
In theory all results should plot between the lines shown for “slow” model (no
inertia forces) and “rapid” model (inertia forces considered). It is evident from Figure
5.52 that this is in fact not the case as a significant number of cases plot below the
“slow” model. Table 5.20 summarises the characteristics of the case studies and their
post failure deformation behaviour in relation to where they plot in Figure 5.52.
For the case studies that plot from about mid range between the “slow” and “rapid”
model up to the “rapid” model generally include cases where strain weakening at or post
the initial failure was considered significant and post failure velocities were in the
higher end of moderate to rapid range. The Khalili et al (1996) “rapid” model is
considered a reasonable upper bound for these cases. It is an upper bound because of
the assumption that the potential energy of the slide mass is consumed entirely as
frictional energy along the surface of rupture in the model (based on the residual
strength). In reality, a portion of the potential energy will be consumed in
disaggregation of the slide mass, and in the initial stages of failure the frictional energy
consumed will be greater than that assumed using residual strength parameters.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.102

Table 5.19: Results of the post-failure deformation analysis using the Khalili et al
(1996) method.
1 ?i – ?f *1 Deform- Velocity
?i *
Slope Type Case FS res *2 Observed ation Category Comment
(º) 3
o
() (m) * *4
Wothorpe B 18.2 0.53 – 0.62 4.2 – 7.6 2 to 3.5 slow Deformation over 3
year period,
translational slide
Cut Slopes in
Barrowden 18.8 0.65 – 0.75 1.0 – 1.1 0.5 to 0.55 slow Rotational slide
Upper Lias
Clay Heyford 22.7 0.53 – 0.61 3.8 – 4.4 2 to 2.3 suspect Compound slide
slow
Seaton 20.1 0.58 – 0.67 2.6 – 2.9 2 to 2.2 suspect Translational slide
slow
Wembley Hill 22.5 0.56 – 0.63 7.9 6 rapid Retained cut, Type 5
Northolt 22.3 0.60 – 0.69 2.4 – 3.6 0.8 to 1.2 slow Type 1, 22 degree
cut.
Sudbury Hill 20.0 0.66 – 0.85 4.1 – 5.7 1.6 to 2.2 slow to
moderate
St. Helier 20.3 0.52 – 0.66 5.8 – 6.3 2 to 2.3 slow to
moderate
Cut Slopes in West Acton B 18.1 0.64 – 0.79 1.5 – 9.1 0.41 to 2.5 slow Limited deformation
London Clay at crest of slope.
Althorne 21.3 0.54 – 0.65 4.5 – 6.7 2 to 3 slow
Kingsbury S1 22.7 0.59 – 0.65 1.7 – 3.5 0.5 to 1.05 slow
Dedham 23.5 0.57 – 0.65 6.1 – 14.7 2.1 to 5 moderate Large deformation at
(toe) toe of slide.
Isle of Sheppey 35.6 0.21 – 0.28 38 12.5 (toe) rapid (?) Steep cut slope of 38
-A degrees.

Carsington 19.1 0.70 – 0.81 8.3 – 9.4 14.3 to 18.1 moderate Slide during
construction
San Luis 17.9 0.70 – 0.87 5.2 – 5.6 16 to 17.3 moderate Slide during
drawdown
Embankment Siburua 26.9 0.65 – 0.73 8.5 – 9.4 1.7 to 1.9 moderate Slide post
construction
Dams
Dam FD2 (1994 23.0 0.80 – 0.83 0.4 0.43 slow Reactivation during
drawdown) drawdown

Old Eildon 22.7 0.64 – 0.78 16.2 13.5 moderate Slide during
(1929 drawdown) drawdown

Rio 8.5 0.70 3.0 0.68 slow to Ductile CH clay


moderate foundation
Fill on Soft St. Alban-A 21.5 0.75 8.1 1.15 rapid Sensitive clay
Ground * 5 foundation.
Portsmouth 14.5 0.27 10.4 3.3 to 3.4 rapid (?) Sensitive clay
foundation.
Road Fill Case RF1 23.5 0.57 – 0.64 7 to 8 3.5 to 4.3 slow Pre-sheared basal
Embankment plane. Slope = 25o
(avg.)
*1 ?i and ?f are the initial and final positions of the centre of gravity (refer Figure 5.51).
*2 FSres is the factor of safety calculated using residual strengths along the surface of rupture.
*3 Deformations are the actual post failure deformations along the surface of rupture.
*4 Velocity categories are to the IUGS (1995) classification system (refer Table 1.1 in Section 1.3).
*5 for the fill on soft ground case studies the post failure deformations given are crest settlements.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.103

1
Cut slopes in Upper Lias clay
0.9 Cut slopes in London clay
Embankment dams
17 ?
15 Fills on soft ground
0.8 20
7 14 Case RF1
2
Residual Factor of Safety .

1 18
0.7 6 19
9 Khalili et al (1996)
16
4 22 12 Rapid model
11 10
0.6
5
3 8
0.5

0.4 ?

0.3
21 13

0.2
Khalili et al (1996)
0.1 Slow model

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
1−θ f θ i
1 - Wothorpe B 5 - Wembley Hill 9 - West Acton B 13 - Sheppey A 17 - Dam FD2 20 - St. Alban A
2 - Barrowden 6 - Northolt 10 - Althorne 14 - Carsington 18 - Old Eildon 21 - Portsmouth
3 - Heyford 7 - Sudbury Hill 11 - Kingsbury S1 15 - San Luis dam 19 - Rio 22 - Case RF1
4 - Seaton 8 - St. Helier 12 - Dedham 16 - Siburua

0.85
Khalili et al (1996)
17 ? "rapid" model

0.8
Khalili et al (1996) 15
Residual Factor of Safety .

"slow" model
7 20 14
0.75
2
1

0.7 9
19
16
6
0.65

4
22 12
11
0.6
10 5
8
3
0.55
0.0 0.1 0.21 − θ f θi 0.3 0.4 0.5

Figure 5.52: Observed deformation versus the calculated residual factor of safety.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.104

Table 5.20: Characteristics of the case study groupings from the post failure
deformation analysis (Figure 5.52)
Case Study Case Characteristics *1
Grouping Nos.
Cases that plot well Cuts in heavily over-consolidated clays of Type 1 and Type 2 slope
2, 3, 4, 6, failure geometry with cut slope angle typically less than 20 to 25
below the “slow”
11 degrees. Post failure velocity in the slow to low end of moderate
model. category and limited post failure deformation, less than 1.5 to 2.5 m.
Cuts in heavily over-consolidated clays of Type 1 and Type 2 slope
failure geometry with cut slope angle typically less than 20 to 25
degrees. Post failure velocity in the slow to low end of moderate
category.
1, 5, 7, 8, Slides in embankment dams of relatively small volume, and of slow
Cases that plot close 9, 10, 12, to moderate post failure velocity and limited deformation (less than 2
to the “slow” model. 16, 17, m). Limited post failure brittleness of slide mass and mechanism.
19, 22 Fills on high plasticity soft clay foundations of low sensitivity. Slow
to moderate post failure velocity.
Road embankment case study (slope = 25 degrees (avg) and Type 2
slope failure geometry) and Type 5 retained cut slope in London clay
(Wembley Hill).
Significant post failure brittleness of slide mass and mechanism for
most slides, and includes:
• Slides in embankment dams of relatively large volume (>
Cases that plot from 13, 14, 300,000 m3). Post failure slide velocity in the moderate to rapid
mid range up to the 15, 18, range.
“rapid” model. 20, 21 • Steep angled Type 2 cuts in heavily over-consolidated high
plasticity clays, with post failure velocity in the rapid range.
• Fills on brittle, highly sensitive, soft clay foundations with peak
post failure velocity in the rapid range.
*1 References to post failure velocity are to the IUGS (1995) classification system.
The embankment dam case studies analysed remained virtually intact on sliding and reached
a peak velocity no greater than the moderate category.

The five case studies that plot well below the “slow” model are all slides in cut slopes of
heavily over-consolidated high plasticity clays of slow post failure velocity and post
failure deformations less than about 1.5 to 2.5 m. In all cases the surface of rupture was
fully formed in cross section as indicated by the development of a back-scarp and
displacement in the toe region of the slide. The reason why these slides plot well below
the “slow” model is likely to be related to insufficient displacement along the surface of
rupture for residual strengths to have been reached and to strain hardening on post
failure shearing for these over-consolidated clays. The typical deformation behaviour
for these “slow” slides in relatively flat cut slopes is for deformation to occur as an
initial failure followed by subsequent reactivations over a period of several years (refer
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.105

Section 5.7.3). If the monitored deformation only captures the initial post failure
deformation of the slide then it will plot below the “slow” model in Figure 5.52.
Of the cut slope case studies that plot close to the “slow” model, the information
from the published literature indicates that the deformation of the slide mass, in most
cases, has accumulated over a period of at least 3 years and is representative of several
reactivations of slide movement. Therefore, the assumption of residual strength on the
surface of rupture is probably a reasonable one for these cases. Case RF1 is also
representative of initial sliding and a number of subsequent reactivations of movement.
It too plots close to the “slow” model.
It is notable that the slide at Wembley Hill (Type 5 slope failure geometry in a
retained cut slope), which travelled at a rapid post failure velocity, plotted close to the
“slow” model. The reason for this may be associated with energy losses and effects on
the lateral margins where shearing occurred through the existing retaining wall.
For the embankment dam case studies, the slides of smaller volume (Siburua and
Dam FD2) and/or slow post-failure velocity (Dam FD2) plotted close to the “slow”
model. The slides of relatively large volume (greater than about 100,000 m3) and of
moderate post-failure velocity (Old Eildon, San Luis and Carsington dams) plotted from
midway between the “slow” and “rapid” model up to “rapid” model. The results for the
embankment dam analyses would suggest the greater inertia of the larger volume slides,
for a given slide velocity, is significant in the type of model used for prediction of travel
distance, but the data is insufficient to confirm this.
For the fill on soft ground case studies, the failures in low plasticity sensitive clay
foundations reached post failure velocities in the rapid range due to the highly strain
weakening properties of the sensitive clays at failure (refer Section 4.3.5.2 for further
details). For these cases the “rapid” model provides a reasonable upper bound for travel
distance prediction.

5.8.2 EMPIRICAL METHODS FOR PREDICTION OF TRAVEL DISTANCE OF INTACT


SLIDES

Empirical analysis has been undertaken on the slides in embankment dams and cut
slopes in heavily over-consolidated soils that were considered to have reached post-
failure velocities greater than about the mid moderate range (greater than about 2000 to
5000 mm/day) and that travelled distances in excess of about 5 m. Case studies where
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.106

the post-failure velocity was in the slow to low end of moderate range were excluded
from the analysis given the limited travel distance of these slides. The slides with Type
5 slope failure geometry were also excluded because of the limited significance of slope
angle on the post failure deformation behaviour.
From the analysis of the post failure deformation behaviour of “rapid” landslides in
soil slopes (Section 3.9 of Chapter 3), reasonable correlations were found for the travel
distance angle with respect to slope angle (both cut slope angle and downslope angle),
slide volume and degree of confinement. Definitions of the terminology are given in
Section 1.3. Figure 5.53 defines the travel distance angle, a, travel distance, L, and
landslide height, H.
For the embankment dam case studies the empirical analysis indicates that
statistically no correlation exists between the ratio H/L (tangent of the travel distance
angle) and slide volume (Figure 5.54). With respect to the tangent of the cut or fill
slope angle (Figure 5.55), statistically a correlation does exist with the ratio H/L
although it is relatively weak given the limited number of cases (R2 = 0.62 for 14 cases).
For the cut slopes in heavily over-consolidated clays the number of cases is insufficient
to undertake a statistical analysis. Combining the two sets of case studies does not
improve the statistical correlation.

Figure 5.53: Schematic profile of slope failure showing travel distance and travel
distance angle

The lack of correlation of H/L with slide volume is to be expected given the significant
effect of the material strength properties (strength at failure and strain weakening on
shearing), slide mechanism and slide geometry on the post failure distance of travel.
The statistical correlation with respect to slope angle is also to be expected due to the
limited distance of travel beyond the initial toe of the slide (5 m up to a maximum of 15
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.107

to 20 m) for the moderate to rapid post-failure slide velocity cases and the co-
dependence between slope angle and kinetic energy of the slide mass for slides in
similar soils, such as for the cut slope failures of Type 2 slope failure geometry in
London clays. This is partly reflected in Figure 5.55 as a trend of increasing distance of
separation with increasing cut slope angle between the point representing the case study
and the line representing H/L = slope angle.
The range of cut slopes angles over which limited deformations at slow post failure
velocity were observed in London clays is also shown in Figure 5.55. For cut slope
angles less than about 25 to 30 degrees limited post failure deformations were generally
observed indicating that at these slopes the travel distance angle would be closely
coincident with the cut slope angle.
For the embankment dam case studies a similar correlation is likely to exist;
however, it is complicated by variation in material properties, zoning geometry and
slide mechanism.

0.7

Embankment dams
0.6
Cuts in over-consolidated CH clays

0.5
Ratio H/L

0.4

0.3

Embankment Dams
0.2
no correlation

0.1

0.0
1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Volume (cu.m.)

Figure 5.54: H/L versus the slide volume for embankment dams and cut slopes of
moderate to rapid post-failure velocity.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.108

0.7
Embankment dams
H/L = slope angle
0.6 Cuts in over-consolidated CH clays
Type 1 slope failure geometry with
"rapid" post failure velocity
0.5
Ratio H/L

0.4 Coincidence of H/L


and cut slope angle
for slides in London
0.3 clays of slow post
failure velocity

0.2 Embankment Dams


2
Y = 0.458X + 0.105, R = 0.62
6 3
0.1 Vol. Range = 1000 to 1.2 x 10 m

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Tangent of Fill and Cut Slope Angle
Figure 5.55: H/L versus the tangent of the fill or cut slope angle for embankment dams
and cut slopes of moderate to rapid post-failure velocity.

5.9 CONCLUSIONS AND GUIDELINES F OR P REDICTION OF P OST-


F AILURE DEFORMATION B EHAVIOUR OF INTACT SLIDES IN SOIL
SLOPES
The findings from the case study analysis of the mechanics of failure and mechanics of
post failure deformation behaviour of slides in embankment dams (excluding hydraulic
fills) and in cut slopes in heavily over-consolidated high plasticity clays are summarised
in Section 5.9.1. Guidelines for the prediction of post-failure travel distance of slides in
these slope and material types are presented in Section 5.9.2.

5.9.1 SUMMARY OF FINDINGS FROM THE CASE STUDY ANALYSIS

5.9.1.1 Case Studies


Fifty-four case studies for failures in embankment dams (excluding hydraulic fill dams)
have been analysed. They have been sub-divided into failures during construction,
failures that occur on drawdown and failures that occur post construction in the
downstream shoulder.
In most cases the slide mass remained relatively intact during sliding. Post failure
travel distances ranged from less than 1 m up to about 30 m, and in most cases was less
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.109

than 5 to 10 m. Post failure velocities were generally in the slow to moderate IUGS
(1995) categories (i.e. centimetres per day to metres per hour) with a limited number in
the rapid category. Slide volumes ranged from about 5000 cubic metres up to more than
1 million cubic metres.
Of the failures in cut slopes in heavily over-consolidated high plasticity clays, thirty-
seven case studies were analysed, all from cuts in London clay and Upper Lias clay in
the UK.

5.9.1.2 Embankment Types Susceptible to Slope Instability


The findings of the statistical analysis by Foster (1999) of slope instability incidents in
large dam embankments and further analysis of the Foster database (refer Section 5.2.1)
as well as the case study analysis, indicates that:
• Overall, slope instability is:
− Much more likely in homogeneous earthfill, puddle core and concrete core-wall
earthfill embankments;
− Much less likely in rockfill and zoned earthfill embankments;
− Much more likely in embankments constructed of medium to high plasticity
clays and silts;
− Where the foundation is involved in the slide, slope instability is much more
likely in embankments founded on medium to high plasticity fine-grained soils
(mainly clays), and much less likely in embankments founded on bedrock or
sandy and gravelly soils;
− Where the slide is through the embankment only, slope instability is much less
likely in zoned or core-wall embankments with dominantly sandy or gravelly
materials or rockfill in the outer zones, provided the outer zone is of substantial
thickness;
• For failures during construction, the foundation is involved in a high proportion of
slope instability incidents, about 75%;
• For failures post construction, either during drawdown or in the downstream
shoulder:
− A high proportion of incidents occur in homogeneous embankments (94% of
incidents in homogeneous embankments occur post construction);
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.110

− A high proportion of failures in the downstream shoulder occur within earthfill


embankments (homogeneous, earthfill with filters and earthfill with rock toe),
mainly in embankments without embankment filters or more permeable zones in
the downstream shoulder.

These findings point to the dominance of material type, predominantly medium to


high plasticity clays, in the likelihood of slope instability in embankment dams, and this
could be more generally extended to encompass fills. This is exclusive of fills
susceptible to flow liquefaction, such as the failures in hydraulically placed sandy fills
(refer Chapter 3).

5.9.1.3 Mechanics of Failure


The stress conditions within the slope and the material strength properties are the two
important factors controlling failure in soil slopes. The stress conditions within the
slope are affected by the slope geometry, pore water pressure conditions and coefficient
of earth pressure at rest, Ko.
The effect of defects (bedding, jointing, fissuring, etc.) and layering/zoning on the
strength of the soil mass is dependent on the strength of the defect (or layer) in relation
to that if the intact soil. Under the stress conditions within a slope the shear stresses (as
a percentage of shear stress capacity of the soil or defect) will be higher where the
strength of the defect (or of a weak layer) is lower than that of the intact soil. The
orientation of defects (and weak layers) is an important factor in the stability of the
slope and in the mechanics of failure.

(a) Cut Slopes in Heavily Over-Consolidated High Plasticity Clays


Progressive failure is significant in the mechanics of failure of slides in cut slopes in
heavily over-consolidated high plasticity clays. Localised failure initiates in regions of
high shear stress, typically in the toe region of the cut, and progressively spreads into
the slope (Potts et al 1997). The stress relief process (as shown by Potts et al) is
important in the continuation of progressive failure in these slopes due to the gradual
change in effective stress condition as pore water pressures equilibrate, and dilation and
softening of the soil mass occurs. The permeability of the soil mass is an important
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.111

factor in the stress relief process as it directly affects the time for equilibration of pore
water pressures.
Stress relief and progressive failure, amongst other factors, results in a gradual
decrease in the overall stability of slope with time (Picarelli 2000). The trigger for
failures in cut slopes in heavily over-consolidated high plasticity clays is often rainfall
or seasonal wetter periods. But, for steeper slopes it can be the processes of stress relief
and progressive failure.

(b) Embankment Dams


For failures in embankment dams during construction, the trigger is often increasing
stress conditions due to embankment raising. The mechanics of failure can, and in a
number of cases is, simply undrained failure under the stress conditions imposed.
Most of the cases of failure during construction involve low undrained strength
layers in either the foundation or the earthfill (wet placed zones or wet placed layers
within a zone). The very high percentage of clay soils involved in failures during
construction is often related to their low undrained strength, high construction pore
water pressures and slow dissipation of pore water pressures in these low permeability
soil types. Progressive failure can be significant in the mechanics of failure for slides
during construction, as was the case in the slide at Carsington dam.
For failures during drawdown, it is the drawdown itself that triggers the slide due to
the change in stress conditions within the slope and reduction in support of the water
load on the upstream face. Embankments with clayey soils in the upstream shoulder
predominate in slides during drawdown due to their low permeability and high pore
water pressures retained in the slope under the drawdown condition. Other factors
though are significant in failures during drawdown:
• The change in effective stress conditions within the slope due to wetting up and
saturation of partially saturated soils.
• The change in material strength properties, mainly clay soils, due to swelling on
saturation.

An important observation of the slides during drawdown is that failure is often not
triggered by the largest and/or fastest drawdown in the history of reservoir operation.
This is due in part to the above factors, but more significantly due to progressive strain
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.112

weakening in undrained loading of over-consolidated low permeability soils, mainly


fine-grained clayey soils.
For failures in the downstream shoulder that occur post construction, seepage and
saturation within the embankment from the impounded water storage is the trigger for
failure in most cases. Changes in effective stress conditions within the slope occur as a
result of saturation of partially saturated soils, rising pore water pressures and seepage
pressures. Material strength properties are also affected by the change in moisture
content and swelling. These failures dominantly occur in fine-grained soils, and once
again material permeability is a significant issue. Rainfall and earthquake have also
triggered slides in the downstream shoulder post construction.
An important issue with the seepage triggered slides is the potential for a breach
condition. Most of these failures (i.e. seepage triggered slides) initiated when the
reservoir was close to full supply level or had reached its historically highest level.

5.9.1.4 Mechanics of Post Failure Deformation Behaviour


The mechanics of post failure deformation is complex. It is complicated by, amongst
other factors, strain and stress dependent material strength properties, time dependent
material properties, slope geometry, orientation of the surface of rupture and the
mechanics of failure. Therefore, simple empirical criteria for assessment of travel
distance and slide velocity in these slopes, such as cut or fill slope angle, are inadequate.
For most failures in embankment dams (excluding hydraulic fill dams) and cuts in
heavily over-consolidated high plasticity clays, the slide remained virtually intact during
sliding. Travel distances were generally relatively limited, less than 5 to 10 m in most
cases but can be up to 25 to 30 m, and post failure velocities typically ranged from the
slow to rapid IUGS (1995) velocity categories, or from metres/year to metres/minute.
The potential energy of the slide mass at failure is defined by geometrical
constraints including the slope geometry and orientation of the surface of rupture. The
amount of this potential energy (if fully realised by the slide) that is consumed as
frictional energy, and the remainder that is available for kinetic energy and energy for
disaggregation of the slide mass is dependent on the material properties and slide
mechanism. Materials that are significantly strain weakening from their strength at
failure and slide mechanisms that incorporate brittleness on slide release, have the
potential for conversion of a significant proportion of the potential energy into kinetic
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.113

energy for acceleration of the slide mass and achievement of greater travel distances at
higher velocities.
In summary, the factors affecting the post failure deformation behaviour of slides in
embankment dams and in cuts in heavily over-consolidated high plasticity clays were:
• Potential for material strain weakening. Significant strain weakening at or post the
initial failure was observed for:
− Near normally consolidated high plasticity clays due to particle orientation and
pore water pressure generation on shearing in undrained loading;
− Over-consolidated mainly clay type soils where progressive strain weakening on
shearing after initial failure (at strengths between peak and critical state)
occurred in either drained or undrained loading;
− Structured soils that are susceptible to large loss in undrained strength on
shearing and consequent increase in pore water pressure as load is transferred
onto the pore fluid (i.e. static liquefaction);
• Brittleness associated with the slide mechanics, such as due to internal brittleness or
toe buttressing;
• Brittleness associated with slide release or reduction of restraining forces on the
lateral margins;
• Slope failure geometry. The slide mass for slides of Type 1 slope failure geometry
generally has greater potential energy than for Type 2 and in turn Type 5 slope
failure geometries;
• Orientation of the surface of rupture. This will influence the potential energy
realised by the slide mass (e.g. for a slide of Type 1 slope failure geometry, where
the slide remains intact in the source area, has not realised its potential energy
compared to a slide that fully evacuates the source area) and to some extent the
distribution of the potential energy to frictional energy.

The analysis of post failure deformation behaviour of the case studies (Sections 5.6
and 5.7), particularly for the failures in embankment dams, attempted to differentiate
those slides where post failure brittleness of the slide mechanics was significant from
those where brittleness was not greatly significant. The assessment of significance of
brittleness was based on the likelihood of acceleration of the slide mass due to strain
weakening of the slide mechanism or materials as discussed above. Where it was not
significant, post failure deformations were likely to be limited and at slow post failure
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.114

velocity. Where it (brittleness of the slide mechanics) was significant, post failure
travel distances were generally greater and the post failure velocity of the slide mass
higher, typically of moderate to rapid velocities (i.e. metres/hour to metres/minute or
faster).
The following sub-section summarise the findings from the case study analysis on
factors, both general and more specific, that influenced the post failure deformation
behaviour of the various sub-groups of slides analysed. General guidelines for
estimation of slide velocity are based on the analysis and are given for the purpose of
allowing slide velocity estimation for improved prediction of the post failure travel
distance (Sections 5.8.1 and 5.9.2) and for other risk assessment purposes. They should
be used with caution.

5.9.1.5 Post Failure Deformation Behaviour of Slides in Embankment Dams


The general guidelines for estimation of post failure slide velocity in embankment dams
are divided into the three groups; failures during construction, failures during drawdown
and post construction failures in the downstream shoulder. The factors summarised
below are for landslides where the post failure velocity was in the moderate to rapid
IUGS (1995) classes or faster (i.e. metres/hour to metres/minute or faster).

(a) Failures during construction


For most landslides in embankment dams during construction the post failure
deformation of the slide was driven by out of balance forces as a result of fill placement.
In most cases the slide decelerated on cessation of fill placement and the travel distance
was of limited extent and generally at a slow post failure velocity.
Of importance though is the group of landslides that accelerated after cessation of
filling due to progressive strain weakening of the slide mass. For a number of these
slides, material strain weakening of near normally high plasticity clays in the foundation
or earthfill due to particle orientation was sufficiently significant for acceleration of the
slide mass for these relatively large volume failures. Progressive failure in the over-
consolidated high plasticity clays was significant in the failure at Carsington dam and
static liquefaction due to contraction on shearing of structured soils was likely to have
been significant in the rapid failure at Marshall Creek dam.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.115

For several of these slides of higher post failure velocity, brittleness of the slide
mechanism due to release or weakening on the margins, passive toe resistance and/or
internal shearing was significant. In these cases the persistence of the weak zone in the
foundation or earthfill was typically of broad extent both in cross section and in long
section. Therefore, even though the resistance on the lateral margins was significant,
strain weakening on the surface of rupture is likely to have resulted in high shear
stresses on the margins sufficient to cause either formation of a shear surface on the
margins or large lateral extension of the slide, and result in moderate to rapid slide
velocities and relatively large travel distances.

(b) Failures in the upstream slope during drawdown


The material type, slope failure geometry, upstream slope angle and the drawdown itself
were significant influences on the post failure deformation behaviour of slides in the
upstream shoulder triggered by drawdown. Table 5.21 provides a summary of the how
these factors affect the post failure velocity. For about half of the failure case studies,
post failure travel distances were of limited extent and estimated peak velocities were in
the slow to low end of moderate category.
Significant factors in the moderate to rapid failures during drawdown were:
• Progressive failure in over-consolidated medium to high plasticity clay earthfills and
high plasticity foundations, where strain weakening occurred on shearing after the
initial slope failure occurred;
• Brittleness in the slide mechanics at and post failure, such as from concrete facing
elements; and
• The slope failure geometry and upstream slope angle. For slides of Type 1 slope
failure geometry, the toe of the slide mass broke up and travelled at a “rapid”
velocity down the face of the upstream slope where the upstream slope was equal to
or greater than 2H to 1V (about 27 degrees).

(c) Failures in the downstream slope post construction


The slide volumes of failures in the downstream shoulder post construction were much
smaller than for the failures during construction and during drawdown (Table 5.11).
The factors contributing to higher post failure velocity of the slide mass (i.e. post failure
velocities generally in the mid moderate to rapid range) were:
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.116

• Material type. Slides involving high plasticity clays tended to show higher post
failure velocity, more so for slides of Type 2 slope failure geometry than Type 5;
• Progressive failure. Where progressive failure was significant in the mechanics of
failure and post failure deformation, velocities were likely to be higher;
• Trigger to failure is seepage related. Those failures where pore water pressure and
seepage were considered to trigger the failure tended to travel at higher post failure
velocities, most probably due to softening of the slide mass from seepage infiltrating
the dilating slide mass during post failure deformation.

Table 5.21: Factors affecting the peak post-failure velocity for slides in embankment
dams during drawdown.
Material Type/s IUGS (1995) Velocity Categories
(along the surface of Slow Moderate Rapid
rupture)
High plasticity silts - Slopes less than or -
and clays equal to 3H to 1V
Low to medium Slopes of 2 - 3H to 1V Slopes of about 2H to Type 1 failure and/or
plasticity silts and 1V up to about 1.5H to brittleness to slide
clays 1V. mechanism (e.g.
concrete facing).
Slopes greater than or
equal to 2H to 1V.
Sandy and Gravelly Slopes of 2 - 3H to 1V. - Extremely rapid
soils with significant Very rapid sustained drawdown, Type 1
fines content drawdown to trigger failure, Slope steeper
failure (> 500 than or equal to 2H to
mm/day). 1V.

5.9.1.6 Post Failure Deformation Behaviour of Slides in Cut Slopes in Heavily Over-
Consolidated High Plasticity Clays
The potential for progressive strain weakening of the material once failure had occurred,
slope geometry and slope failure geometry were significant factors in the post failure
travel distance and velocity of slides from failures in cut slopes of heavily over-
consolidated high plasticity clays.
For the failures in cut slopes in London clay, a rapid post failure velocity was
observed, in most cases, where the timing of the failure occurred within days to less
than ten years after excavation for slides of Type 1 and Type 2 slope failure geometry,
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.117

and within 15 years for Type 5 slope failure geometry (based on a limited number of
case studies). For these slides this relatively early time of failure (for cuts in London
clay) may reflect a higher average shear strength at failure on the surface of rupture and
greater potential for strain weakening to residual strength post failure. Therefore,
greater slide velocity is generally observed where there is greater potential for material
strain weakening from initial failure.
The post failure deformation behaviour for the cuts slopes showed a distinctive
change in behaviour with increasing cut slope angle for slides of Type 1 and Type 2
slope failure geometry (Figure 5.45 and Figure 5.46). In summary, rapid post failure
velocity is likely for:
• Slides of Type 1 slope failure geometry in cut slopes steeper than about 22 degrees.
For these cases the toe region of the slide mass will potentially break up and travel
at a moderate to rapid velocity for relatively large distances on the lower portion of
the cut slope. With increasing cut slope angle above about 22 degrees the volume of
material likely to evacuate the source area is likely to increase.
• Slides of Type 2 slope failure geometry in cut slopes steeper than about 30 to 35
degrees.

5.9.2 GUIDELINES FOR PREDICTION OF POST-FAILURE DEFORMATION

The post-failure deformation analysis undertaken (Section 5.8) included simplistic


empirical type analyses and limit equilibrium related analyses based on the Khalili et al
(1996) method for intact sliding.
The results of the empirical analyses, related to slide volume and cut (or fill) slope
angle, are presented in Section 5.8.2. No correlation was found with respect to slide
volume. The analyses with respect to cut or fill slope angle indicated that a statistical
correlation does exist, however, it is relatively weak (Figure 5.55). Better correlations
were obtained with respect to slope angle where the case studies in the analysed group
consisted of similar material type, slope failure geometry and/or mechanism (e.g. Figure
5.40 for drawdown related failures). Overall, the empirical methods are considered to
be of limited use, and useful only for a broad based estimate of travel distance. It is
better to rely simply on the general performance (Section 5.9.1).
The limit equilibrium modelling and Khalili et al (1996) method proved to be a
useful method for prediction of post-failure travel distance. The results of the analysis
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.118

are presented in Figure 5.52 (in Section 5.8.1). Guidelines for deformation analysis
using the Khalili et al (1996) model are summarised in Table 5.22. The guidelines are
mainly for those slides that remain basically intact or coherent on sliding and are not
susceptible to static liquefaction or other phenomenon that can lead to “rapid” sliding.
Excluded from the Khalili et al (1996) analysis were slides in hydraulic fill embankment
dams and slides of Type 1 slope failure geometry in relatively steep slopes where the
toe region of the slide mass broke up and travelled “rapidly” down the cut or fill slope
were.
For the embankment dam case studies, reference in Table 5.22 is made to slide
volume and potential peak post-failure slide velocity. The volume categories are
somewhat arbitrary given the limited number of cases analysed, although the modelling
did indicate the larger volume slides of moderate velocity approached the “rapid” model
compared with smaller volume slides of similar post-failure velocity. General
guidelines for estimation of peak post-failure velocity are summarised in Section
5.9.1.4.

Table 5.22: Guidelines for post-failure deformation prediction using the Khalili et al
(1996) models.
Slope Type Khalili et al (1996) “Slow” Khalili et al (1996) “Rapid”
Model Model
Cut slopes in heavily Type 1 slope failure geometry at Type 2 slope failure geometry at
over-consolidated clays cut slopes less than about 22 cut slopes greater than about 25 to
degrees. 30 degrees.
Type 2 slope failure geometry at
cut slopes less than about 25 to 30
degrees.
Type 5 retained cut slopes.
Fills on Soft Ground Fills on soft clay foundations of Fills on brittle, highly sensitive
low sensitivity. clay foundatio ns.
Embankment Dams Large volume (> 100,000 m3) Large volume (> 100,000 m3)
(excluding hydraulic slides with potential peak slides with potential peak
fills) velocities in the slow range. velocities in the moderate to rapid
Smaller volume (< 100,000 m3) range.
with potential for peak velocities Smaller volume (< 100,000 m3)
in the slow to moderate range. with potential for peak velocities
in the rapid range.
Chapter 5: Landslides in Embankment Dams and Cuts in HOC Clays Page 5.119

The Khalili et al (1996) “slow” model often over-estimates the amount of initial
deformation for the materials and slope conditions identified in Table 5.22 (refer Figure
5.52 in Section 5.8.1), and it can therefore be regarded as a reasonable upper bound.
The reason for the over-estimation by the “slow” model for the cut slope case studies
analysed was considered to be related to the mode of post failure deformation of these
slides (generally deformation on initial failure followed by reactivation in subsequent
wet periods), and the likelihood that residual strength is not reached following the initial
failure. Only the slide in the retained cut slope in London clay at Wembley Hill
appeared to contradict this general trend. A similar post failure mode of deformation
was observed for several slides in embankment dams triggered by drawdown (e.g. Dam
FD2 and Bear Gulch dam) and the road fill embankment failure (Case RF1).
For those case studies under the Khalili et al (1996) “rapid” model, post failure
deformations generally occurred within a single event (i.e. no subsequent reactivation).
For the smaller volume slides (including cut slopes and fills on soft ground) the post
failure deformation of the slide usually occurred within a matter of minutes to hours at a
rapid post failure velocity. For the larger volume embankment dam case studies, the
post failure deformation of the slide took from days to months, but typically involved a
phase of large acceleration.
These slides (under the “rapid” model) tended to plot below the post failure
deformation estimated by the “rapid” model. The reason for this is considered to be in
the formulation of the “rapid” model, which assumes that all available potential energy
of the slide mass is resisted by frictional forces along the surface of rupture
approximated by the residual soil strength. The model does not take into consideration
energy losses associated with disaggregation of the slide mass, internal shearing or
energy used in deformation to reach residual strength conditions. Therefore, the “rapid”
model can be considered as an upper bound estimate of the post failure deformation of
the intact slides for the slide categories in Table 5.22.
Chapter 6: Deformation Behaviour of Rockfill Page i

TABLE OF CONTENTS

6.0 THE DEFORMATION BEHAVIOUR OF ROCKFILL.......... 6.1


6.1 OUTLINE OF THIS CHAPTER .............................................................................6.1
6.2 DEFINITIONS AND TERMINOLOGY ....................................................................6.2
6.2.1 Unconfined Compressive Strength of Rock............................................... 6.2
6.2.2 Rockfill Placement and Compaction......................................................... 6.3
6.2.3 Rockfill Moduli..........................................................................................6.4
6.2.4 Zoning of Main Rockfill in Concrete Face Rockfill Dams ........................ 6.7
6.3 LITERATURE R EVIEW OF ROCKFILL DEFORMATION ........................................ 6.9
6.3.1 Historical Summary of Rockfill Usage in Embankment Design................ 6.9
6.3.2 Deformation Properties of Rockfill ........................................................... 6.9
6.3.3 Predictive Methods for Rockfill Deformation ......................................... 6.14
6.4 ANALYSIS OF THE DEFORMATION BEHAVIOUR OF ROCKFILL IN CONCRETE
FACE ROCKFILL DAMS .................................................................................. 6.27
6.4.1 Case Study Database............................................................................... 6.28
6.4.2 Deformation During Construction .......................................................... 6.28
6.4.3 Deformation of the Face Slab of CFRD on First Filling ........................ 6.44
6.4.4 Post Construction Crest Settlement......................................................... 6.53
6.5 DISCUSSION AND RECOMMENDED METHODS FOR PREDICTION .....................6.71
6.5.1 Guidelines on Deformation Prediction During Construction................. 6.71
6.5.2 Guidelines on Deformation Prediction During First Filling .................. 6.75
6.5.3 Guidelines on Deformation Prediction Post-Construction.....................6.76
6.6 CONCLUSIONS ............................................................................................... 6.77

LIST OF TABLES
Table 6.1: Classification of unconfined compressive strength or rock (AS 1726-
1993) ....................................................................................................................... 6.2
Table 6.2: Historical summary of rockfill usage in embankment design (Galloway
1939; Cooke 1984; Cooke 1993). ......................................................................... 6.10
Table 6.3: Parameters for deformation prediction (Soydemir and Kjærnsli 1979)...... 6.25
Chapter 6: Deformation Behaviour of Rockfill Page ii

Table 6.4: Rates of post-construction crest settlement of dumped and compacted


rockfills in CFRDs (Sherard and Cooke 1987) ..................................................... 6.26
Table 6.5: Summary of embankment and rockfill properties for CFRD case
studies.................................................................................................................... 6.29
Table 6.6: Assessment of cross-valley influence on arching for case studies
analysed................................................................................................................. 6.38
Table 6.7: Stress conditions representative of the data sets from Figure 6.21............. 6.43
Table 6.8: Summary of crest settlement during the period of first filling.................... 6.62
Table 6.9: Values of the coefficient, m, in the strain rate – time power function
(Equation 6.5)........................................................................................................ 6.63
Table 6.10: Estimates of long-term crest settlement rates for dumped rockfills.......... 6.67
Table 6.11: Location of internal post-construction vertical settlement in CFRD ........ 6.70
Table 6.12: Approximate stress reduction factors to account for valley shape............ 6.73

LIST OF FIGURES
Figure 6.1: Determination of rockfill moduli, (a) modulus during construction,
Erc, and (b) modulus during reservoir filling, Erf (Fitzpatrick et al 1985)............... 6.5
Figure 6.2: Typical design of dumped rockfill CFRD (adapted from ICOLD
(1989b))................................................................................................................... 6.7
Figure 6.3: Typical zoning of main rockfill in current CFRD design practice for
construction with sound quarried rockfill (adapted from Cooke (1997); zoning
designators after Fell et al (1992)) .......................................................................... 6.8
Figure 6.4: Compression curves for dry and wet states, and collapse compression
from dry to wet state for Pyramid gravel in laboratory oedometer test (Nobari
and Duncan 1972a) ............................................................................................... 6.13
Figure 6.5: Settlement versus time curves for laboratory oedometer tests on
rockfill (Sowers et al 1965)................................................................................... 6.14
Figure 6.6: Typical stress-strain relationship of rockfill from a triaxial
compression test (Mori and Pinto 1988) ............................................................... 6.17
Figure 6.7: Finite element analysis of Foz do Areia CFRD (Saboya and Byrne
1993), (a) model and (b) stress paths during construction and first filling. .......... 6.18
Figure 6.8: Deformation modulus during construction (Erc) versus void ratio
(Pinto and Marques Filho 1998) ........................................................................... 6.20
Figure 6.9: Ratio of deformation modulus on first filling to during construction
(Erf /Erc) versus valley shape factor (Pinto and Marques Filho 1998)................... 6.20
Chapter 6: Deformation Behaviour of Rockfill Page iii

Figure 6.10: Deformation modulus during construction (Erc) for Hydro Tasmania
CFRDs (Giudici et al 2000) .................................................................................. 6.21
Figure 6.11: Results of 3-dimensional finite element analysis of CFRD, to give
vertical displacement during construction versus valley shape (Giudici et al
2000) ..................................................................................................................... 6.22
Figure 6.12: Settlement rate analysis of Cedar Creek dam; (a) determination of
time to and (b) settlement rate versus time (Parkin 1977)..................................... 6.24
Figure 6.13: Post-construction crest settlement of membrane faced compacted
rockfill dams (Clements 1984). ............................................................................. 6.26
Figure 6.14: Settlement index versus time for well-compacted rockfills (adapted
from Public Works Department NSW 1990) ........................................................ 6.27
Figure 6.15: Stress-strain relationship for rockfills observed during construction...... 6.33
Figure 6.16: Secant modulus versus vertical stress from monitoring during
construction........................................................................................................... 6.33
Figure 6.17: Tangent modulus versus vertical stress from monitoring during
construction........................................................................................................... 6.35
Figure 6.18: Idealised model for two-dimensional finite difference analysis of
cross-valley influence............................................................................................ 6.36
Figure 6.19: Results of two-dimensional finite difference analysis of the effect of
cross-valley shape on vertical stresses in the dam. (a), (c) and (e) represent
construction in 5 m lifts (to 100 m) and (b), (d) and (f) construction in a single
100 m lift. .............................................................................................................. 6.38
Figure 6.20: Indicators of cross valley arching effects from variations in the secant
modulus with vertical stress. ................................................................................. 6.39
Figure 6.21: Representative secant modulus (mostly Zone 3A rockfill) at end of
construction (Erc) versus D80 from average grading of the rockfill....................... 6.42
Figure 6.22: Erf / Erc ratio versus embankment height ................................................. 6.45
Figure 6.23: Stress paths during construction and first filling for nominal 100 m
embankment with 1.3H to 1V upstream slope angle; (a) monitoring point
locations, (b) stress paths for points normal to face slab at 30% of the
embankment height. .............................................................................................. 6.49
Figure 6.24: Stress paths during construction and first filling for nominal 100 m
embankment with 1.55H to 1V upstream slope angle; (a) monitoring point
locations, (b) stress paths for points normal to face slab at 30% of the
embankment height. .............................................................................................. 6.49
Figure 6.25: Face slab deflection during first filling (4/2/71 to 25/4/71) of Cethana
dam (Fitzpatrick et al 1973). ................................................................................. 6.50
Chapter 6: Deformation Behaviour of Rockfill Page iv

Figure 6.26: Face slab deformation during first filling of Aguamilpa dam (Mori
1999) ..................................................................................................................... 6.51
Figure 6.27: Face slab deformation during first filling of Ita dam (Sobrinho et al
2000) ..................................................................................................................... 6.51
Figure 6.28: Scotts Peak dam; embankment zoning and location of face cracks
(courtesy of Hydro Tasmania)............................................................................... 6.52
Figure 6.29: Mackintosh dam (a) embankment section (courtesy of Hydro
Tasmania) and (b) face slab deformation on first filling (Knoop and Lack
1985) ..................................................................................................................... 6.52
Figure 6.30: Examples of derivation of zero time for post-construction settlement.... 6.56
Figure 6.31: Post-construction crest settlement versus time for dumped rockfill
CFRD, to at end of main rockfill construction. ..................................................... 6.57
Figure 6.32: (a) and (b) Post-construction crest settlement versus time for CFRD
constructed of compacted rockfills of medium to high intact strength, to at end
of main rockfill construction. ................................................................................ 6.58
Figure 6.33: (a) and (b) Post-construction crest settlement versus time for CFRDs
constructed of well-compacted rockfills of very high intact strength and of
well-compacted gravels, to at end of main rockfill construction. .......................... 6.59
Figure 6.34: Post-construction crest settlement versus time for dumped rockfill
CFRD, to at end of first filling. .............................................................................. 6.60
Figure 6.35: Post-construction crest settlement versus time for CFRD constructed
of compacted rockfills of medium to high intact strength, to at end of first
filling. .................................................................................................................... 6.60
Figure 6.36: Post-construction crest settlement versus time for CFRDs constructed
of well-compacted rockfills of very high intact strength and of well-
compacted gravels, to at end of first filling. .......................................................... 6.61
Figure 6.37: Post construction crest settlement of Bastyan dam.................................. 6.65
Figure 6.38: Long-term crest settlement rate (as a percentage of embankment
height per log cycle of time) versus embankment height for compacted
rockfills ................................................................................................................. 6.66
Figure 6.39: Crest settlement attributable to first filling (excluding time dependent
effects)................................................................................................................... 6.68
Chapter 6: Deformation Behaviour of Rockfill Page 6.1

6.0 THE DEFORMATION BEHAVIOUR OF ROCKFILL

6.1 OUTLINE OF THIS CHAPTER


Rockfill, including other coarse granular materials such as gravels and granular
colluvium, have since the late 19th century been an integral part of embankment dam
construction. Recent and current dam projects incorporating rockfill as a key
component in embankment design include concrete faced rockfill dams (CFRD) now
reaching heights in excess of 200 m and earth and rockfill dams (mostly central core
earth and rockfill dams) reaching heights of 250 to 300 m. The key attributes of rockfill
are its high shear strength, allowing for stable embankment slopes up to 1.3 to 1.4H to
1V (horizontal to vertical), and its low compressibility when placed in a well-compacted
condition.
The stability of rockfill embankments is evidenced by the lack of historical records
of slope instability associated with embankments incorporating rockfill zones. The
statistical data on slope instability incidents in large dams (Foster et al. 2000),
summarised in Section 5.2.1, indicates that of 124 incidents in embankment dams only
7 cases incorporated rockfill zoning. Of the failure case studies of embankment dams
analysed in Chapter 5, rockfill was incorporated in the embankment design in 5 of the
54 case studies. Only in 1 of these 5 cases did slope instability involve shear through
the outer rockfill (it was a zoned earth and rockfill dam with very thin outer rockfill
shoulders); in the other four cases the surface of rupture preferentially passed through
the foundation.
This chapter presents the results of the research on the deformation properties of
rockfill. Section 6.3 presents an historical summary of the usage of rockfill in
embankment design, and review of the literature on the properties of rockfill affecting
its deformation behaviour and methods for prediction of rockfill deformation (during
and post construction). Analysis of the deformation behaviour of 36 well-monitored
rockfill dams (mostly concrete face rockfill dams (CFRD)) is presented in Section 6.4.
The analysis is sub-divided into; the deformation of rockfill during construction, the
face slab deformation of CFRD on first filling, and the post construction crest settlement
of CFRD.
Guidelines on methods for prediction of the deformation behaviour of rockfill are
presented in Section 6.5 based on the findings from the analysis in Section 6.4.
Chapter 6: Deformation Behaviour of Rockfill Page 6.2

Methods based on the historical performance of case studies are developed for
estimation of the rockfill modulus and the post construction crest settlement of CFRD.

6.2 DEFINITIONS AND TERMINOLOGY


Section 1.3 includes a summary of the terminology used to describe the classification of
rockfill placement and the unconfined compressive strength of intact rock. Additional
discussion is presented on these two areas as well as rockfill moduli and embankment
zoning in CFRD in the following sub-sections

6.2.1 UNCONFINED COMPRESSIVE STRENGTH OF ROCK

The classification system from Australian Standard AS 1726-1993 is used to classify


the unconfined compressive strength (UCS) of intact rock used for rockfill. The
descriptors and the UCS range they represent are given in Table 6.1.

Table 6.1: Classification of unconfined compressive strength or rock (AS 1726-1993)


Strength Descriptor UCS Range (MPa)
Extremely High > 240
Very High 70 to 240
High 20 to 70
Medium 6 to 20

For the purposes of the deformation analysis of rockfill, the strength of intact rock has
been simplified into two classes, very high strength and medium to high strength. It is
recognised that the medium to high strength grouping (UCS of 6MPa to 70 MPa) is very
broad, but they have been grouped together for several reasons:
• The database consists of very few cases of medium strength rockfills and the
analysis indicated they could be grouped with the high strength rockfills without
introducing significant uncertainty into the statistical correlations presented;
• On handling, placement and compaction of rockfills sourced from medium to high
UCS rock, particle size distributions generally show that significant breakdown
occurs as reflected in the high percent passing 19 mm and the generally high Cu
(coefficient of uniformity, D60/D10). It is recognised that there is likely to be “grey
Chapter 6: Deformation Behaviour of Rockfill Page 6.3

zone” rather than a division based only on UCS strength, as the amount of
breakdown will be affected by the rock type, type and weight of roller, compaction
procedure and strength of cementing of the rock.

Extremely high strength rock has been used as rockfill in embankment construction
for a few case studies and these have been included with those of very high strength.

6.2.2 ROCKFILL PLACEMENT AND COMPACTION

The method of placement of rockfill has a significant influence on its compressibility


during construction and its deformation behaviour post construction. The definitions by
Cooke (1984, 1993) for dumped and compacted rockfill have been used as a basis for
categorisation of the method of placement. The definitions used are:
• Compacted Rockfill – rockfill placed in layers up to 2 m thickness (generally 0.9 to
2.0 m thick) and compacted by smooth drum vibrating roller (SDVR). Accepted
practice is typically 4 to 6 passes of a minimum 10 tonne (possibly up to 15 tonne)
deadweight vibrating roller, with variation in layer thickness, added water and
number of passes depending on the quality and type of the rockfill, amount of fines
and location within the embankment. Three classifications for compacted rockfill
have been used:
− Well-compacted – layer thickness typically less than about 1.0 m (depending on
the compressive strength of the intact rock) and compacted with a minimum of
four passes of a 10 to 15 tonne deadweight SDVR.
− Reasonable Compaction – layer thickness typically 1.5 to 2.0 m and compacted
with typically four passes of a 10 tonne SDVR.
− Reasonably to Well Compacted - layer thickness typically 1.2 to 1.6 m
(depending on the compressive strength of the intact rock) and compacted with
typically 4 to 6 passes of a 10 to 15 tonne SDVR.
• Rockfill not formally compacted or “poorly compacted”. Several methods of
rockfill placement have been included under the definition “poorly compacted”,
these include:
− Dumped rockfill – rockfill placed in lifts ranging from several to tens of metres
thickness, with or without sluicing, and without formal compaction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.4

− Rockfill placed in lifts less than about 2 to 3 m thickness and not formally
compacted (i.e. without the use of rollers for compaction). Specified track
rolling by bulldozer or other plant, or rockfill indicated as being trafficking by
trucks or other haulage equipment has been classified under “not formally
compacted”.
− Rockfill placed in lifts greater than 2 to 3 m and formally compacted. For these
rockfills the lift thickness is considered too great for compaction to have any
significant influence at depth.

Watering is an important component for placement of rockfills, particularly in cases


where the compressive strength of the rock used in the rockfill is susceptible to
reduction on wetting, breaks down under the action of the roller, or if the rockfill
contains large quantities of fines. Cooke (1993) comments that watering is not overly
important for compaction of very high strength rockfills that are not susceptible to
weakening on wetting. However, these rockfills can still show collapse type settlements
on wetting.
For dumped rockfills, sluicing had a significant influence on the deformation
behaviour of the rockfill as evidenced by the large collapse deformations of dry dumped
or poorly sluiced rockfills when wetted (Cogswell dam (Baumann 1958), Strawberry
and Dix River dams (Howson 1939)). Terminology used to define the level of sluicing
of dumped rockfill has been broadly defined into three classes; dry dumped, poorly
sluiced and well sluiced. A well-sluiced rockfill is described (Steele and Cooke 1958)
as rockfill sluiced at a water ratio of 2 - 3 to 1 (water to embankment fill volume) using
high-pressure jets that are directed on the tipped material.

6.2.3 ROCKFILL MODULI

Fitzpatrick et al (1985) defined two moduli for assessment of the deformation behaviour
of rockfill (Figure 6.1), the rockfill modulus during construction, Erc, and the rockfill
modulus on first filling, Erf, calculated from Equations 6.1 and 6.2. Both moduli are
used extensively throughout this chapter.

E rc = γ H d 1 / δ s (6.1)
Chapter 6: Deformation Behaviour of Rockfill Page 6.5

E rf = γ w h d 2 / δ n (6.2)

where Erc and Erf are in MPa, ? = unit weight of the rockfill in kN/m3, ?w = unit weight
of water in kN/m3, ds = settlement of layer of thickness d1 due to the construction of the
dam to a thickness H above that layer; dn = face slab deflection at depth h from the
reservoir surface, and d2 is measured normal to the face slab as shown. H, h, d1 and d2
are all measured in meters, and ds and dn are measured in millimeters.

Figure 6.1: Determination of rockfill moduli, (a) modulus during construction, Erc, and
(b) modulus during reservoir filling, E rf (Fitzpatrick et al 1985)

Erc is determined from the settlement of hydrostatic settlement gauges (HSG) installed
during construction, generally from under the embankment centreline. The modulus is a
secant modulus and, for HSGs under the embankment centreline, very closely
approximates the confined modulus. The method by Fitzpatrick et al (1985), shown in
Figure 6.1a, does not take into consideration the stress distribution effect of
embankment shape.
Where possible the calculation of Erc for the case studies analysed has taken into
account the distribution in applied vertical stress with depth due to embankment shape.
For case studies where detailed data was available on internal settlements (from HSGs)
as construction proceeds it has been possible to calculate the rockfill modulus (both
secant and tangent modulus) with increasing vertical stress, allowing for the effects of
embankment shape. The effect of embankment shape has been allowed for by using the
stress intensity factors for elastic solutions by Poulos et al (1972) and vertical stresses
calculated from finite element analysis. In summary, the method used to calculate the
vertical stress and the rockfill moduli between say two HSGs or between a HSG and the
foundation, was:
Chapter 6: Deformation Behaviour of Rockfill Page 6.6

• Take the mid point between the HSGs as representative of the vertical stress of the
layer of rockfill (the HSGs were usually a vertical distance of 10 to 20 m apart).
• On placement of the upper HSG, estimate the initial vertical stress at the mid point
of the layer from elastic solutions. This is the stress at the zero settlement reading of
the upper HSG.
• For subsequent fill heights above the upper HSG determine the vertical stress (at the
layer mid-point) appropriate to each fill height from elastic solutions.
• The secant modulus (or Erc) at each fill height can then be determined by dividing
the change in vertical stress (from the initial vertical stress) by the vertical strain in
the rockfill zone of interest (i.e. the settlement between the HSGs) appropriate to the
change in vertical stress.

The elastic solutions by Poulos et al (1972) are for an embankment with slopes of 30
degrees (1.75H to 1V, horizontal to vertical) constructed of materials with a Poisson’s
Ratio of 0.3 on a rigid foundation. From finite difference analysis it was found that
these solutions reasonably approximated the vertical stresses for embankments with
slopes as steep as 1.3H to 1V without introducing significant errors. For HSGs located
off centreline, the effect of the offset was taken into consideration in the estimation of
vertical stress. Only HSGs from within the central half of the embankment (in most
cases on or very close to the embankment centreline) and in the lower 40 to 60% of the
embankment were used for estimation of the rockfill moduli during construction. A
Poisson’s Ratio of 0.3 is considered a reasonable estimate for rockfill.
In comparison to the vertical stresses under the embankment centreline estimated by
the Fitzpatrick et al (1985) method, equivalent stresses are obtained up to 60 to 70
percent of the embankment height. For construction above this height (i.e. on
construction of the upper 30 to 40%) the embankment shape influences the vertical
stresses and the Fitzpatrick et al 1985 method will over-estimate the actual stresses and
under-estimate the rockfill moduli.
Erf is calculated from the deformation of the face slab normal to the upstream face
on first filling as shown in Figure 6.1b. Fitzpatrick et al (1985) comment that the
modulus calculated is only an indicative value of the modulus of the rockfill on first
filling, but that the face slab deformations calculated using this method reasonably (and
simply) approximate those from finite element analysis. The method is used very
broadly throughout the concrete face rockfill dam (CFRD) community. In reality, Erf is
Chapter 6: Deformation Behaviour of Rockfill Page 6.7

not a true modulus of the rockfill, rather it is an artefact of the method of calculation and
an index of performance useful for the estimation of face slab deflection. It should only
be used for this purpose. This is discussed further in Section 6.4.3.

6.2.4 ZONING OF MAIN ROCKFILL IN CONCRETE FACE ROCKFILL DAMS

Zoning descriptors used for the main rockfill zones varies widely throughout the CFRD
community. A uniform system of classification of the main rockfill zones has been
applied to the CFRD case studies analysed. It broadly follows the classification systems
of Cooke and Sherard (1987), ICOLD (1989b) and Fell et al (1992).
For CFRDs of dumped rockfill construction (typical of the CFRD design of the
1920’s to 1960’s) the typical embankment design incorporates the main body of
dumped rockfill (designated Zone 3) and an upstream zone of derrick placed rockfill
supporting the facing membrane (Figure 6.2).

Figure 6.2: Typical design of dumped rockfill CFRD (adapted from ICOLD (1989b))

Following the resurgence of CFRD usage in dam design with the change to compaction
of rockfill from about the mid to late 1960’s, current design practice is dependent on a
number of factors including the minimisation of deformation of the face slab on first
filling and control of any leaks which may occur through the face slab or seepage
through the foundation. For rockfills sourced from quarried sound rock the design of
the main rockfill (Figure 6.3) typically incorporates two zones; Zone 3A and Zone 3B.
Zone 3A, the main support zone for the upstream face, is typically placed in 1 m layers
and well compacted to achieve a high modulus for minimisation of face slab
Chapter 6: Deformation Behaviour of Rockfill Page 6.8

deformation. For Zone 3B, the downstream rockfill zone, a high modulus is not as
critical as for Zone 3A and the rockfill is typically coarser and placed (and compacted)
in 1.5 to 2 m thick layers.

Figure 6.3: Typical zoning of main rockfill in current CFRD design practice for
construction with sound quarried rockfill (adapted from Cooke (1997); zoning
designators after Fell et al (1992))

The central variable zone in Figure 6.3 represents variations in the sizing of Zones 3A
and 3B in current design practice around the world. Hydro Tasmania (formerly Hydro-
Electric Commission, Tasmania) typically design with a dominant Zone 3A forming 75
to 100% of the main rockfill (i.e. Zone 3B is only present in the outer portion of the
downstream shoulder). The typical design of large CFRDs in Brazil (of sound quarried
rockfill) consists of Zone 3A comprising the upstream third of the main rockfill and
Zone 3B the downstream two thirds.
Design of the main rockfill zones becomes more complex when the rockfill is to
consist of potentially low permeability materials, such as dirty gravels or medium to
high strength quarried rockfills that breakdown significantly on compaction. Drainage
zones behind the face slab and above the foundation are generally incorporated into the
design for control of seepage. The zoning classification of the main rockfill used for
designs other than typical is based on similar principles as for the typical design, with
Zone 3A designated the main upstream zone and Zone 3B, if used, the main
downstream zone. To be designated Zone 3A the rockfill zone must comprise at least
25% of the main rockfill.
In the case of Aguamilpa dam (refer Figure 6.26 in Section 6.4.3) the main rockfill
incorporated three zones designated Zone 3A for the upstream zone of compacted
Chapter 6: Deformation Behaviour of Rockfill Page 6.9

gravels, Zone 3T for the central/downstream rockfill zone and Zone 3B for the outer
downstream rockfill zone.
For details on current design features for the concrete face slab, plinth slab, jointing
details, upstream facing zones, upstream toe zoning and non-typical CFRD designs for
low strength and low permeability rockfills refer to Cooke and Sherard (1987), ICOLD
(1989b), Fell et al (1992), Cooke (1999) and Marulanda and Pinto (2000).

6.3 LITERATURE R EVIEW OF ROCKFILL D EFORMATION

6.3.1 HISTORICAL SUMMARY OF ROCKFILL USAGE IN EMBANKMENT DESIGN

An historical summary of the use of rockfill in embankment design and construction


(Galloway 1939; Cooke 1984; Cooke 1993) is given in Table 6.2.
From the 1930’s to the mid 1960’s the height of construction of CFRDs was limited
to about 80 to 100 m due to extensive face slab cracking and high leakage rates as a
result of excessive deformation of the dumped and sluiced rockfill. In terms of stability
though (Cooke 1984) the embankment performance was excellent. The transition from
dumped and sluiced to compacted rockfill in embankment design (Cooke 1984, 1993)
occurred in the late 1950’s to 1960’s and resulted in proliferation in the use of CFRD
from the late 1960’s.

6.3.2 DEFORMATION PROPERTIES OF ROCKFILL

From the mid 1930’s (Galloway 1939; Morris 1939) it was recognised that a significant
component of rockfill deformation was associated with particle breakage at point
contacts under increasing loads during construction and on reservoir filling. Peterson
(1939) commented on the “critical importance” of sluicing in obtaining settlements
during the construction period. Terzaghi (1960) surmised that the particle breakage that
occurred under increasing stress conditions resulted in the rearrangement in the granular
structure to a more stable position, giving the large deformations during construction
and impounding and the subsequent significant reduction in the rate of deformation post
initial impounding.
Chapter 6: Deformation Behaviour of Rockfill Page 6.10

Table 6.2: Historical summary of rockfill usage in embankment design (Galloway 1939;
Cooke 1984; Cooke 1993).
Approximate Method of Placement and Comments
Time Period Characteristics of Rockfill
Concrete Face Rockfill Dams
Mid to late Dumped rockfill with timber facing Early embankments constructed with timber facing.
1800’s to early Typically of very steep slopes (up to 0.5 to 0.75H to
1V). First usage of concrete facing in the 1890’s.
1900’s
Height limited to about 25 m.

1920’s to Main rockfill zone dumped in high Rockfill typically sound and not subject to
1930’s lifts (up to 20 to 50 m) and sluiced, disintegration. Dam heights reaching 80 to 100 m.
although the sluicing was relatively For high dams, cracking of the facing slab and joint
openings resulted in high leakage rates (2700 l/sec
ineffective. A hand or derrick placed
Dix River, 3600 l/sec Cogswell, 570 l/sec Salt
rockfill zone was used upstream.
Springs).

Late 1930’s to High pressure sluicing used for the Cracking of face slab, particularly at the perimeter
1960’s main rockfill zone. Rockfill still very joint, and high leakage rates a significant issue with
coarse. higher dams (3100 l/sec at Wishon, 1300 l/sec at
Courtright).

From late Rockfill placed in 1 to 2 m lifts, Significant reduction in post-construction


1960’s watered and compacted. Reduction in deformations due to low compressibility of
particle size. Usage of gravels and compacted rockfill. Significant reduction in leakage
rates; maximum rates typically less than 50 to 100
lower strength rock.
l/sec. Continued improvement in plinth design and
facing details to reduce cracking and leakage.

Earth and Rockfill Dams


1900 to 1930 Dumped rockfill Use of concrete cores with dumped rockfill shoulders
at angle of repose. Limited use of earth cores. Dam
heights up to 50 to 70m.

1930’s to Earth core (sloping and central) wi th Use of earth cores significant from the 1940’s due to
1960’s dumped rockfill shoulders. the difficulties with leakage of CFRD. Increasing
dam heights up to 150 m.

From 1960’s Use of compacted rockfill. Typically Improvements in compaction techniques. Early dams
placed in 1 to 2 m lifts, watered and compacted in relatively thick layers with small
compacted with rollers. rollers. Gradual increase in roller size and reduction
in layer thickness reduced the compressibility of the
rockfill. Significant increase in dam heights in the
mid to late 1970’s, up to 250 to 300 m.

With the increased usage of rockfill in embankment construction of high dams from the
1960’s came the use of large scale laboratory testing (mainly oedometer, and triaxial
and plane strain shear testing) to better understand the strength and deformation
Chapter 6: Deformation Behaviour of Rockfill Page 6.11

properties of rockfill. The testing confirmed the significance of particle breakage on the
deformation behaviour of rockfill under increasing applied stresses and on saturation.
In summary, it was found that the deformation (and modulus) of rockfill was
predominantly affected by:
• The compactive effort in placement of the rockfill. Increased modulus was
observed with increased compactive effort (Marsal 1973).
• The applied stress level. Increased particle breakage and decreasing modulus were
observed with increasing deviatoric stress levels in triaxial compression tests
(Marsal 1973; Marachi et al 1969). In oedometer tests relatively high moduli were
observed for compacted rockfill samples up to normal stresses in the order of 800 to
1000 kPa, thereafter the modulus was observed to decrease with increasing normal
stress (Marsal 1973).
• The stress path. Significantly higher modulus was observed on un-loading and re-
loading at stress levels less than previously experienced by the rockfill as shown in
Figure 6.6 (Mori and Pinto 1988).
• The particle shape and grading of the rockfill. Greater deformation (and lower
modulus) was observed for angular (in comparison to rounded), more uniformly
graded (i.e. lower coefficient of uniformity, Cu) and coarser rockfills (of similar Cu
but higher maximum particle size) (Marachi et al 1969; Marsal 1973; Bowling
1981).
• Intact rock strength. Reduced modulus, greater deformation and reduced strength
were observed for weaker strength rockfills (Marsal 1973).

A significant aspect of rockfill deformation behaviour is its propensity to collapse


on wetting. Probably the most infamous case of collapse settlement was observed
during the construction of Cogswell dam (CFRD) in 1933 (Baumann 1958). The
rockfill was dumped in high lifts, up to 46 m, without sluicing. Following a very heavy
rainstorm in December 1933 the embankment crest was observed to settle up to 4.6 m
(or 5.4%) and bulging on the mid to lower slopes caused significant damage to the
facing. High pressure sluicing after the rainstorm resulted in further crest settlements of
up to 1.7 m. Similar, but not as dramatic, collapse deformations during construction on
flooding of poorly sluiced rockfills were observed at Strawberry and Dix River dams
(Howson 1939). Terzaghi (1960) attributed the deformation on saturation to a loss in
Chapter 6: Deformation Behaviour of Rockfill Page 6.12

strength of the rockfill, mainly in the outer surface of the particles and commented that
it was more likely to occur in weathered rockfills.
The results of laboratory testing investigating the collapse deformation behaviour of
rockfill indicate:
• The stress-strain curve of rockfill in a dry state is of lower compressibility than for
similar rockfill (similar grading and density) in a saturated state (Figure 6.4).
• Collapse deformation on wetting occurs for dry rockfill when the stress state (in
stress-strain space) is above the normal compression line for wetted or saturated
rockfill (Alonso and Oldecop 2000).
• On wetting, the collapse strain is equivalent to the difference in strain (at a given
confining stress) between the dry and wetted states (Figure 6.4) (Nobari and Duncan
1972a; Alonso and Oldecop 2000).
• The initial water content at placement has a significant influence on the amount of
collapse deformation on wetting (Nobari and Duncan 1972a), the higher the
moisture content at placement the lesser the collapse deformation on wetting.
• There is a time delay between flooding of the sample and collapse deformation
(Marsal 1973; Alonso and Oldecop 2000). Martin (1970), as reported by Justo
(1991), using water and organic liquids for sample flooding, found that the amount
of collapse deformation and time for collapse deformation to occur was dependent
on the liquid used for saturation of the sample.
• Collapse deformations of similar magnitude to that occurring on flooding were
obtained by imposing 100% relative humidity on the rockfill sample (Alonso and
Oldecop 2000), indicating that flooding or wetting of the voids between the rock
particles was not required for collapse deformation.

Alonso and Oldecop (2000) concluded that the collapse deformation occurred within
the individual rock particles, and that the amount of deformation was controlled by the
imposed stresses and initial moisture content within the outer exposed voids or pores of
the individual rock particles. They hypothesised that the collapse mechanism was
associated with crack propagation of the rock pores, and that the rate of propagation was
significantly affected by the total suction in the rock pores. They further concluded that
“any situation leading to a change in moisture content in the rock pores is enough to
cause collapse deformation”, which is consistent with the observation of collapse
deformations induced by flooding, such as on reservoir filling, or rainfall.
Chapter 6: Deformation Behaviour of Rockfill Page 6.13

This assessment seems to assume that the reduction in compressive strength on


saturation observed for many rocks, which leads to failure at the highly stressed point
contacts in an angular rockfill mass, is entirely dependent on suction effects in the outer
pores of the rock pieces at the highly stressed point contacts. This seems unlikely.

Figure 6.4: Compression curves for dry and wet states, and collapse compression from
dry to wet state for Pyramid gravel in laboratory oedometer test (Nobari and Duncan
1972a)

From laboratory oedometer tests on rockfill (Sowers et al 1965; Marsal 1973; Alonso
and Oldecop 2000) it is observed that the deformation on application of an increase in
applied stress occurred as a large, almost instantaneous component followed by a much
smaller time-dependent strain component. Sowers et al (1965) commented that this
secondary time-dependent component of strain approximated a straight line when
plotted against the log of time (Figure 6.5). They further commented that the log rate of
strain increased with increasing applied stress for dry rockfill samples and on saturation
(post the collapse deformation). Alonso and Oldecop (2000) observed similar findings
and hypothesised that this time dependent strain was due to the on-going process of
crack propagation and particle breakage.
Chapter 6: Deformation Behaviour of Rockfill Page 6.14

Figure 6.5: Settlement versus time curves for laboratory oedometer tests on rockfill
(Sowers et al 1965)

6.3.3 PREDICTIVE METHODS FOR ROCKFILL DEFORMATION

For CFRD and the rockfill zones of earth and rockfill dams the components of rockfill
deformation are deformations that occur due to load application (during construction
and impoundment), on saturation or wetting and on-going time dependent or creep type
deformations. Predictive methods are typically divided into the three components
deformation during construction, on first filling and long-term post construction (or post
first filling).
Most predictive methods cover the deformation behaviour of one or two of these
components. Finite element analyses have been used (Saboya and Byrne 1993;
Kovacevic 1994; amongst others) for analysis of rockfill deformation during
construction and on first filling treating the events sequentially, and depending on the
embankment type, consider the deformation on saturation only during first filling.
Empirical methods, usually based on historical performance of embankments, are
typically available for prediction of deformation during construction and post-
construction. The deformation on first filling and on wetting or saturation is implicitly
incorporated in the post-construction deformation component for these methods. In the
case of CFRD, methods are available that specifically consider face slab deformation
Chapter 6: Deformation Behaviour of Rockfill Page 6.15

during first filling, and in these cases, the rockfill is assumed to remain un-wetted or
unsaturated on impoundment.
From field observations the significant factors affecting the deformation behaviour
of rockfill in embankment dams (mainly from CFRD) are reported to include:
• Degree of compaction of the rockfill;
• Applied stress conditions and stress path (Mori and Pinto 1988);
• Particle shape and size distribution;
• Intact strength of the rock used as rockfill;
• Wetting or saturation of the rockfill causing collapse deformation;
• Time dependent or creep type deformations.

Pinto and Marques Filho (1998) and Giudici et al (2000) consider that the geometric
shape of the valley has a significant influence on the deformation behaviour due to cross
valley arching, resulting in a reduction in vertical stresses within the embankment.
Deformation moduli calculated using Equation 6.1, which ignores the three-dimensional
effect of valley shape on vertical stress, will over-estimate the vertical stress and
therefore over-estimate the moduli.
Available predictive methods incorporate several of these factors as discussed in the
following sub-sections. Finite element analyses may incorporate most of these factors
in the stress-strain relationships on which the constitutive models are based, but are
generally not used to model the time-dependent deformations.

6.3.3.1 Finite Element Analyses


Duncan (1996) and Kovacevic (1994) are recent references covering the state of the art
in finite element analyses applicable to the deformation behaviour of embankments
(mainly zoned earth and rockfill dams) during construction and on first filling. They
discuss the methods of analysis, their limitations, available constitutive models of the
stress-strain relationship and areas of uncertainty. As Duncan points out, most analyses
from the literature are Class C1 (Lambe 1973), i.e. after the event, which may account
for the generally good agreement between predicted and observed deformation
behaviour.
Duncan (1996) comments that the choice of stress-strain constitutive model used in
the analysis is a balance between simplicity and accuracy, and that the choice of
Chapter 6: Deformation Behaviour of Rockfill Page 6.16

constitutive model will depend on the purpose of the modelling. He commented that if
the purpose of the analysis is to analyse stresses and/or the trend of deformations then
more simplistic models could be suitable, depending on the relative stiffness between
different embankment materials. More accurate modelling of deformations requires a
more complex stress-strain model that more realistically approximates the real
behaviour of the rockfill and earthfill.
Several other important aspects of finite element analysis with respect to modelling
dam embankments (more specifically toward embankments constructed of rockfill)
raised by Duncan (1996) and Kovacevic (1994) include:
• The importance of modelling the construction as a series of incremental layers
• The type of stress-strain relationship used for modelling materials. Linear elastic
models are not suitable for accurate modelling of deformation behaviour due to the
non-linear stress-strain relationship of rockfill (Figure 6.6). Hyperbolic or multi-
linear elastic models (determined from laboratory oedometer tests) can reasonably
model deformations during construction given that the stress paths in oedometer
testing are similar to the stress paths under the centreline of the embankment and
that plastic deformations under the shear stress conditions are generally not
significant for rockfill embankments.
• A limitation of hyperbolic and multi-linear elastic models is that the model
parameters are typically derived from triaxial compression or oedometer laboratory
tests and are therefore limited by the narrow range of stress paths covered in
comparison to the broader range of stress paths imposed in the field. Kovacevic
(1994) highlighted the limitations of multi-linear elastic models to accurately model
the deformation of the face slab of concrete face rockfill dams (CFRD) on reservoir
impoundment. This would also apply to zones in other embankment types (e.g. the
downstream shoulder in central core earth and rockfill dams) where the stress path
on impoundment is markedly different to that in the triaxial compression or
oedometer test.
• Kovacevic (1994) found that elasto-plastic models that model pre-peak plasticity
were more suited for modelling the deformation behaviour of the upstream face of
CFRD during reservoir impoundment due to the ability of these constitutive models
to more realistically account for the rockfill deformation under the stress path
conditions imposed (Figure 6.7).
Chapter 6: Deformation Behaviour of Rockfill Page 6.17

• A significant factor of uncertainty associated with Class A type predictions based on


laboratory testing is the difference in material properties between the laboratory test
results and those in the field due to limitations on the maximum particle size that
can be tested and variations in stiffness due to differences in material strength,
compacted density and moisture content.

Figure 6.6: Typical stress-strain relationship of rockfill from a triaxial compression test
(Mori and Pinto 1988)

A further problem with laboratory test results is the limitation of representing the
layering and segregation that occurs within any single layer of rockfill. The upper part
of the layer is broken down to a greater degree under the action of the roller than the
lower part of the layer and segregation occurs during placement. As a result of the field
compaction process this layering effect results in density, modulus and grading
variations within a single layer.
An important component of the modelling of embankment dams is the consideration
of collapse compression of susceptible rockfills and earthfills on wetting. The effects of
collapse compression are most noted for the upstream shoulder on initial impoundment,
but collapse compression has also been observed in the downstream shoulder following
wetting due to rainfall, leakage or tail-water impoundment. Incorporating collapse
compression into constitutive models adds further complexity and greater uncertainty in
the estimation of material parameters between laboratory and field conditions due to the
dependency of collapse compression on compaction moisture content, compacted
density, applied stress conditions and material properties. Justo (1991) and Naylor et al
(1989) propose methods for incorporation of collapse compression of rockfill into
Chapter 6: Deformation Behaviour of Rockfill Page 6.18

constitutive models. The analysis of Beliche Dam, a central core earth and rockfill dam,
by Naylor et al (1997) is an example where collapse compression of the upstream
rockfill was considered in the modelling.

Figure 6.7: Finite element analysis of Foz do Areia CFRD (Saboya and Byrne 1993),
(a) model and (b) stress paths during construction and first filling.

6.3.3.2 Empirical Predictive Methods of Deformation Behaviour During


Construction and on First Filling
As previously discussed, Class A predictions of deformation behaviour of rockfill
embankments during construction and first filling using finite element analysis are often
based on stress-strain relationships of rockfills from laboratory testing or from
empirically derived properties. Laboratory testing requires specialist laboratory
equipment, particularly for rockfill, to accommodate particle size distributions that are
Chapter 6: Deformation Behaviour of Rockfill Page 6.19

at least somewhat representative, although still scaled down, of the placed rockfill.
Available facilities to undertake this testing are also limited and the testing is expensive.
Depending on the proposed embankment design, construction materials and compaction
specifications justification for laboratory testing of rockfill may not be warranted.
Several methods are available for estimation of the rockfill modulus from historical
records of performance of CFRD for well-compacted rockfills. Poulos et al (1972)
provide a simplistic procedure, based on linear elastic analysis, for estimation of the
deformation during construction from non-dimensional factors.
Most of the methods based on historical performance serve to highlight the
significant factors associated with the field deformation behaviour, although, Pinto and
Marques Filho (1998) proposed a simple method to provide a “rough estimate” of the
modulus for estimation of the face slab deflection of CFRD. From historical records of
the performance of CFRDs constructed typically of well-compacted angular rockfills
they conclude that void ratio and valley shape (Figure 6.8) are the dominant influences
on the calculated two-dimensional deformation modulus during construction for narrow
valleys (note that valley shape affects the vertical stress, and will therefore affect the
calculated Erc using the method by Fitzpatrick et al (1985), refer Equation 6.1 and
Figure 6.1a). They define a narrow valley as having a valley shape factor (SF) of less
than 3.5 (SF = A/H2, where A = area of the upstream face slab measured normal with the
upstream face in square metres and H = embankment height in metres).
For wide valleys (shape factor, SF, greater than 4), Pinto and Marques Filho (1998)
comment that the effect of the valley shape can be disregarded and the deformation
modulus (ranging from 30 to 60 MPa) is dependent on the void ratio. For well-
compacted rockfills they comment that the void ratio is dependent on the rockfill
gradation, with higher void ratios representative of poorly graded rockfills and lower
void ratios of well-graded rockfills. For CFRD constructed of rounded gravels Pinto
and Marques Filho (1998) indicate that the deformation modulus during construction is
very high, in the order of 200 MPa.
Pinto and Marques Filho (1998) recommended that the maximum face slab
deformation, D, on first filling be estimated from Equation 6.3, based on empirical
correlation of the ratio of the deformation modulus on reservoir filling (Erf) to during
construction (Erc) with the valley shape factor (Figure 6.9).
Chapter 6: Deformation Behaviour of Rockfill Page 6.20

 0.003   H 2 
D =   ⋅ 
0.21(1+ A H 2 )   E  (6.3)
e   rc 

where D = face slab deflection in metres, A = area of the upstream face slab in m2, H =
embankment height in metres and Erc = deformation modulus during construction in
MPa. Pinto and Marques Filho (1998) comment that the deformation estimate is a
rough estimate, but that it is sufficient for design of CFRDs.

Figure 6.8: Deformation modulus during construction (Erc) versus void ratio (Pinto and
Marques Filho 1998)

Figure 6.9: Ratio of deformation modulus on first filling to during construction (Erf /Erc)
versus valley shape factor (Pinto and Marques Filho 1998)
Chapter 6: Deformation Behaviour of Rockfill Page 6.21

Guidici et al (2000), based on field observations of well-compacted rockfill from Hydro


Tasmania CFRDs, also considered that valley shape had a significant effect on vertical
stress and therefore the deformation moduli calculated during construction from
Equation 6.1 (Figure 6.10), and on that calculated on first filling, Erf, from Equation 6.2
(refer Figure 6.1b). Higher deformation moduli during construction (Erc) were
generally calculated for narrower valleys (as expected given that the vertical stresses
will be over-estimated by ignoring the 3-dimensional effect of valley shape). They
commented that the scatter is associated with other factors affecting the deformation
moduli including intact rock strength and stress levels (as a function of embankment
height). Guidici et al (2000) further investigated the effects of valley shape by
undertaking 3-dimensional finite element analysis of an idealised CFRD and
commented that the results confirm the influence of valley shape on deformation
behaviour during construction (Figure 6.11) and on first filling due to arching effects.

Figure 6.10: Deformation modulus during construction (Erc) for Hydro Tasmania
CFRDs (Giudici et al 2000)
Chapter 6: Deformation Behaviour of Rockfill Page 6.22

Figure 6.11: Results of 3-dimensional finite element analysis of CFRD, to give vertical
displacement during construction versus valley shape (Giudici et al 2000)

6.3.3.3 Predictive Methods for Deformation Behaviour Post Construction

For CFRD and other rockfill embankments, post-construction time dependent


deformations are observed for more than 30 years after construction (Sherard and Cooke
1987), typically at a gradually reducing strain rate with time. Empirical predictive
methods of the post-construction deformation from the literature are typically based on
historical deformation curves for similar embankment types or empirical relationships
derived from historical performance of deformation behaviour. Several of the methods
incorporate the deformation during first filling whilst others only consider the prediction
of deformation post first filling.
Sowers et al (1965) proposed a logarithmic relationship between crest settlement
and time (Equation 6.4) to describe the post construction crest settlement for “rockfill”
dams. It was derived from a database of 14 rockfill dams (mix of concrete face rockfill,
central core earth and rockfill, and sloping core rockfill dams) with rockfill ranging
from sluiced and compacted to dumped and poorly sluiced. Sowers et al found that the
coefficient α was dependent on the method of placement of rockfill and ranged from 0.2
%/log cycle of time for compacted and well sluiced rockfills up to 0.7 %/log cycle of
time. They commented that a straight line could not be used to approximate the post
construction settlement at Dix River dam where the rockfill was dumped and poorly
sluiced. It should be noted that at the time of publication of this paper rockfill
placement procedures were in the transitional phase from dumped to compacted and
Chapter 6: Deformation Behaviour of Rockfill Page 6.23

therefore the level of compaction was dissimilar to what would currently be considered
well-compacted.

∆H = α (log t 2 − log t1 ) (6.4)

where ∆H = crest settlement as a percentage of dam height, t1 and t2 are time in months
from the date when construction was half completed, and α = slope of the crest
settlement-time curve (in units of settlement as a percentage of dam height per log cycle
of time, time in months).
Parkin (1977) commented that a rate analysis, based on the power law relationship
derived from rate process theory (Equation 6.5), allows for more accurate prediction of
post construction settlement of rockfill and is a more powerful tool for evaluation of
performance because it encapsulates the fundamentals of creep deformation. For a
power coefficient of -1 (i.e. m = 1), integration of Equation 6.5 converts to the
logarithmic form of Equation 6.4, and Parkin comments that for many purposes this
approximation may be adequate for predictive purposes. He further indicates that
problems with interpretation or prediction of settlement versus time records using
Equation 6.4 is in the incorrect establishment of the initial time (to) or where m is not
equal to 1 from the rate process equation, which will result in curvature in the settlement
versus log time plot. Figure 6.12a shows the estimation of time to (from the plot of the
inverse of settlement rate versus time) and Figure 6.12b the settlement rate versus time
(t – to) plot with slope of m for Cedar Creek dam.

.
ε = a( t − t o ) − m (6.5)

where ε& = settlement rate in percent/month, a = constant, to = initial or origin of time, t


= time in months after to and m = slope of the settlement rate versus time plot (Figure
6.12b).
Soydemir and Kjærnsli (1979) analysed the post construction deformation behaviour
of 23 membrane faced rockfill dams (mostly CFRD), and proposed an empirical
correlation (Equation 6.6) for prediction of deformation with respect to embankment
height. Table 6.3 presents the coefficient values (ß and d) for maximum crest
settlement, crest horizontal displacement and deflection of the membrane normal to the
Chapter 6: Deformation Behaviour of Rockfill Page 6.24

upstream slope for dumped and compacted rockfill on initial impounding and after 10
years service.

Figure 6.12: Settlement rate analysis of Cedar Creek dam; (a) determination of time to
and (b) settlement rate versus time (Parkin 1977)

s = βH δ (6.6)

where s = deformation in metres, H = embankment height in metres, and ß and d are


constants.
Clements (1984) studied the post-construction deformation behaviour of 68 rockfill
dams, comprising membrane faced (dumped and compacted), sloping core and central
core embankments. Clements proposed upper and lower bound limits of crest
settlement and crest displacement for each type of dam differentiating between dumped
Chapter 6: Deformation Behaviour of Rockfill Page 6.25

and compacted rockfill. The individual deformation curves and bounding limits for the
crest settlement of compacted rockfill for the membrane-faced dams are presented in
Figure 6.13. In comparing the predicted deformations of Soydemir and Kjærnsli (1979)
with the data for the case studies he analysed, Clements concluded that the errors were
too large (correlation coefficients of 0.274 to 0.897) to make any reasonable predictions
using their method. He recommended the use of deformation curves of existing dams
(e.g. Figure 6.13) of similar design and construction methods for prediction of
deformation behaviour. The zero time has been taken as the time of the initial
measurement after the end of construction.

Table 6.3: Parameters for deformation prediction (Soydemir and Kjærnsli 1979)
Coefficients for β and δ (Equation 6.6)
Membrane Faced Membrane Faced
Dumped Rockfill Compacted Rockfill
β δ β δ
Maximum crest settlement, sV :
- Initial inpounding 0.0005 1.5 0.0001 1.5
- 10 years service 0.001 1.5 0.0003 1.5
Maximum crest horizontal
displacement, sH:
- Initial inpounding 0.00035 1.5 0.00005 1.5
- 10 years service 0.0006 1.5 0.00015 1.5
Maximum deflection normal to
upstream face, sN:
- Initial inpounding 0.01 2 0.002 2

Pinto and Marques Filho (1985) commented that the methods of historical comparison
proposed by Clements (1984) are useful, however, their application is not so
straightforward as the deformation behaviour is strongly dependent on the initial time of
measurement in relation to the end of construction (they suggested that comparisons
could be based on the time from start of first filling). And, as pointed out by Parkin
(1977), the shape of the deformation versus time curve will vary depending on the
established zero time.
Sherard and Cooke (1987) compared the post-construction crest settlement from
nine CFRDs of dumped and compacted rockfills. They observed that the crest
settlement of dumped rockfill is 5 to 8 times greater than for compacted rockfill and the
Chapter 6: Deformation Behaviour of Rockfill Page 6.26

rate of crest settlement (Table 6.4) for dumped rockfill 10 to 20 times that of compacted
rockfill.

Figure 6.13: Post-construction crest settlement of membrane faced compacted rockfill


dams (Clements 1984).

Table 6.4: Rates of post-construction crest settlement of dumped and compacted


rockfills in CFRDs (Sherard and Cooke 1987)
Approximate Rate of Crest Settlement for 100 m
High CFRD (mm/yr)
Type After 5 yrs After 10 yrs After 30 yrs
Compacted Rockfill 3.5 1.5 0.6
Dumped Rockfill 45 30 10

Public Works Department NSW (1990) presented a plot of settlement index (Equation
6.7) versus time (Figure 6.14) for well-compacted CFRDs, defining bounds for low
compressive strength and high compressive strength rockfill. The initial time is
established after the end of construction. The data points have been updated where data
records have been available, several data points have been added and the rockfill has
been categorised with respect to intact UCS strength and coefficient of uniformity (Cu).
Further discussion on these aspects is discussed in Section 6.4.4.
Chapter 6: Deformation Behaviour of Rockfill Page 6.27

Settlement Index = Crest Settlement / (Dam Height/100)2 (6.7)

where crest settlement is measured in millimetres from initial monitoring after the end
of construction and dam height is in metres.

700 1. Mackintosh
Very high intact strength, Cu<15 Very high intact strength, Cu>15 2. Mangrove Creek
Medium to high intact strength Intact strength not known 3. Little Para
600 4. Kangaroo Creek
1
5. Serpentine
3 5 6. Tullabardine
500 low compressive 7. Winneke
Settlement Index .

strength rockfill 2 8. Foz do Areia


9. Reece
400
10. Cethana
11 11. Kotmale
23 4 12. Alto Anchicaya
300 6
18 7 13. Murchison
22
14. Shiroro
200 15. Bastyan
16. Crotty
8 9 10 17. Golillas
14 15 13
100 18. Aguamilpa
16 12 20
19. Outrades 2
high compressive strength rockfill
21 19 17 20. Fades
0 21. Salvajina
0 5 10 15 20 25 30 22. Xingo
Time (years) 23. White Spur

Figure 6.14: Settlement index versus time for well-compacted rockfills (adapted from
Public Works Department NSW 1990)

6.4 ANALYSIS OF THE D EFORMATION B EHAVIOUR OF ROCKFILL IN


CONCRETE F ACE ROCKFILL DAMS
CRFDs are recognised for their stability, so one can be confident that measured
deformations are not linked to instability. Conventional limit equilibrium analyses are
generally not undertaken for CFRD (Sherard and Cooke 1987) unless potential stability
problems associated with the foundation are identified, such as shear along a weak
bedding seam in the foundation.
The deformation behaviour of rockfill during construction, on first filling and long-
term post-construction (post first filling) has been analysed from the case study database
of thirty-six mainly CFRD. The factors likely to affect the deformation behaviour of
rockfill that have been considered in the analysis include placement methods, UCS of
the rock used as rockfill, particle size distribution (grading and size), particle angularity,
wetting or saturation, embankment geometry and valley geometry.
Chapter 6: Deformation Behaviour of Rockfill Page 6.28

As part of the analysis, current methods (mainly empirical methods) for prediction
of deformation are evaluated and, where appropriate, what are considered improved
methods have been developed.

6.4.1 CASE STUDY DATABASE

The database of case studies (Table 6.5) includes thirty-five CFRDs and one central
core earth and rockfill embankment (El Infiernillo dam). A summary of the
embankment details and rockfill properties of the case studies is presented in Table 6.5
with further details given in Appendix E.
Selection of case studies was limited, as far as practicable, to CFRDs on rock
foundation or those with a limited depth/area of alluvial gravels in the foundation, with
good quality monitoring records and adequate information on the properties of the rock
used in the rockfill. Case studies were targeted so as to assess the potential factors
affecting deformation behaviour, including dumped versus compacted rockfill, and very
high strength versus medium to high strength rockfill. The deformation during
construction at El Infiernillo dam has been included because of the quality of published
internal deformation records of dry placed and poorly compacted rockfill, for which
little information was found in the literature for dumped rockfills in CFRDs.

6.4.2 DEFORMATION DURING CONSTRUCTION

The analysis of rockfill deformation behaviour during construction has been based on
the observed deformation behaviour from the monitoring of internal settlements mostly
from hydrostatic settlement gauges. This is the most frequently used method of
deformation monitoring in CFRDs. For the most part deformations from gauges
installed under the embankment centreline have been used. Unless otherwise stated, the
data is for Zone 3A rockfill or the equivalent zone/s. Internal horizontal movements are
generally monitored less frequently in CFRD and are not considered here.
Chapter 6: Deformation Behaviour of Rockfill Page 6.29

Table 6.5: Summary of embankment and rockfill properties for CFRD case studies
Embankment
Material Parameters / Properties of Rockfill
Dimensions Time For First Filling
2
Dam Name 4 Strength* Dry Layer Erf (MPa) (from end of main
Hgt, H Upstream Zone * Rockfill Particle d max % finer Void Level of Erc (MPa),
n Cu Density
L/H 1 Class Thickness Placement rockfill construction)
(m) Slope Source* Shape (mm) 19 mm 3 Ratio, e Compaction average
Class Auth UCS (MPa) (t/m ) (m)
Aguamilpa 4p 10t SDVR, in
185.5 2.6 1.5 to 1 3A 3B dredged alluvium (Very High) rounded 85 600 34 2.22 0.18 0.6 Good 305 (250 to 330) 770 0.48 to 1.83 years
(Zone 3A) moist condition
8-12p 6-10t SDVR,
Crotty 83 2.9 1.3 to 1 3A 3A Gravels (Pleitocene) (Very High) rounded 70 200 48 2.54 0.20 0.6 Good 375 (113 to 636) 470 0.87 to 2.0 years
watered
gravels, 4p 10t SDVR, 155 (145 to 165), 4 to 5.17 years (June 82 to
Golillas 125 0.9 1.6 to 1 3A 2 (Very High) rounded 125 350 40 2.135 0.24 0.6 Good 250
unprocessed water added arching likely Aug 83)
Salvajina gravels, 4p 10t SDVR, 205 (175 to 260) -
148 2.4 1.5 to 1 3A 2 (Very High) rounded 9.2 400 32 2.24 0.25 0.6 Good 500 0.33 to >0.75 years (1985)
(Zone 3A) unprocessed water added arching likely
Aguamilpa UCS = 180
185.5 2.6 1.5 to 1 3T T ignimbrite angular 30 500 28 2.04 0.24 0.6 4p 10t SDVR Good 104 - -
(Zone 3T) Very High
4p 10t SDVR, 20% 138 (100 to 170), 375 (@ 30 to
Alto Anchicaya 140 1.9 1.4 to 1 3A D hornfels (Very High) angular 18 600 22 2.28 0.294 0.6 Good 0.18 to 0.2 (Oct 74)
water likely arching 40% height)
8p (10t ?) SDVR,
Bastyan 75 5.7 1.3 to 1 3A 3A Rhyolite, SW to FR (Very High) angular 42 600 25 2.20 0.23 1.0 Good 130 (120 to 140) 290 0.86 to 1.05 years
20% water
160 (120 to 210)
4p 10t SDVR, 15%
Cethana 110 1.9 1.3 to 1 3A 3A Quartzite (Very High) angular 23 900 21 2.07 (0.27) 0.9 Good 105 (85 to 120) 300 0.26 to 0.48 years (1971)
water
likely arching
UCS = 80 6p 10t SDVR, 25% -0.5 to 4 years (12/88 to
Chengbing 74.6 4.4 1.3 to 1 3A 3B/3C Tuff lava angular 10.4 1000 - 2.06 0.277 1.0 Good 43 110
Very High water 1993).

3
Foz Do Areia * basalt (max 25% UCS = 235 4p 10t SDVR, 25% 0.5 to 0.9 years (Apr to Aug
160 5.2 1.4 to 1 3A 1B angular 6 600 10 2.12 0.33 0.8 Good 47 (38 to 56) 80 (65 to 92)
(Zone 3A) basaltic breccia) Very High water 80)

3 6p 9t SDVR, 10%
Ita (Zone 3A) * 125 7.0 1.3 to 1 3A E1/E3' basalt (Very High) angular 11 700 12 2.179 0.308 0.8 Good 48 87 (83 to 91) 0.7 to 0.9 years
water
UCS = 25 angular to 4p 10t SDVR, 0 to 1.95 yrs (Sept 69 to Aug
Kangaroo Creek 60 3.0 1.3 to 1 3A 3 schist 310 600 44 2.34 0.201 0.9 to 1.8 Good - 140
Medium to High subangular 100% water 71)
4p 10t SDVR, 15%
Khao Laem UCS < 190 0.6 to 1.9 years (June 84 to
130 7.7 1.4 to 1 3A 3A limestone angular - 900 - - - 1.0 water lower in 70 Good 59 (43 to 79) 130 to 240
(Zone 3A) Very High Nov 85)
m
charnockitic / (High to Very 4p 15t SDVR, 30% 0.54 to 1.04 years (Nov 84
Kotmale 90 6.2 1.4 to 1 3A 3A angular - 700 - 2.20 - 1.0 Good 61 (47 to 87) 145 (135 to 155)
gneissic High) water to Aug 85)
UCS = 8 - 14 angular, 4p 9t SDVR, no 0.58 to 2.83 years (Aug 77
Little Para 53 4.2 1.3 to 1 3A 3A dolomitic siltstone 110 1000 35 2.15 0.223 1.0 Good 21.5 (19.5 to 23.5) -
Medium elongated ? water lower half to Nov 79)
Greywacke, some UCS = 45 angular, 8p 10t SDVR, 10% 45
Mackintosh 75 6.2 1.3 to 1 3A 3A 52 1000 38 2.20 0.24 1.0 Good 63 1.92 to 2.93 years
slate High elongated water (35 to 60)
Mangrove Creek fresh siltstone & UCS = 45 - 64 angular to 4p 10t SDVR, 0.6 to > 15 years (not
80 4.8 1.5 to 1 3A 1B 310 400 27 2.24 0.18 0.45 to 0.6 Good 55 to 60 -
(Zone 3A) sandstone High subangular 7.5% water reached FSL by 1996)
weathered to fresh
Mangrove Creek UCS = 26 - 64 angular to 4p 10t SDVR, dry
80 4.8 1.5 to 1 3B 4 siltstone & 330 450 32 2.06 0.26 0.45 Good 46 (36 to 56) - -
(Zone 3B) High subangular of OMC
sandstone
UCS = 148 8p 10t SDVR, 20% 190
Murchison 94 2.1 1.3 to 1 3A 3A Rhyolite (SW to FR) angular 19 600 22 2.27 0.234 1.0 Good 560 (485 to 640) 1.04 to 1.09 years
Very High water (170 to 205)
Chapter 6: Deformation Behaviour of Rockfill Page 6.30

Table 6.5 (cont.): Summary of embankment and rockfill properties for CFRD case studies (sheet 2 of 3)
Embankment
Material Parameters / Properties
Properties of
of Rockfill
Rockfill
Dimensions Time For First Filling
2
Dam Name 4 Strength* Dry Layer Erf (MPa) (from end of main
Hgt, H Upstream Zone * Rockfill Particle d max % finer Void Level of Erc (MPa),
n Cu Density
L/H 1 Class Thickness Placement rockfill construction)
(m) Slope Source* Shape (mm) 19 mm 3 Ratio, e Compaction average
Class Auth UCS (MPa) (t/m ) (m)
UCS = 80 - 370 4p 10t SDVR, 5 to 1.32 to 1.46 yrs (Apr to June
Reece 122 3.1 1.3 to 1 3A 3A Dolerite angular 10 1000 11 2.287 0.29 1.0 Good 86 (57 to 115) 190 (175 to 205)
Very High 10% water 1986)
Salvajina weak sandstone 6p 10t SDVR,
148 2.4 1.5 to 1 3B 4 (Medium ?) angular ? 45 600 32 2.26 0.21 0.9 Good 62 (likely arching) - -
(Zone 3B) and siltstone water added
UCS = 22 4-6p 10t SDVR, no 59 (Zone 3A) 0.42 to 2.58 years (June 72
Scotts Peak 43 24.8 1.7 to 1 3A 3A argillite angular ? 380 914 38 2.095 0.266 0.915 Good 20.5 (18.5 to 23.5)
Medium to High water 420 (Zone 2B) to Aug 74)
3
Segredo * basalt (<5% basaltic UCS = 235 6p 9t SDVR, 25%
145 5.0 1.3 to 1 3A 1B angular 7.4 - - 2.13 0.37 0.8 Good 55 (arching likely) 175 Approx. 0.6 years
(Zone 3A) breccia) Very High water

angular to 4p 9t SDVR, not 0.24 to 2.94 years (Dec 71


Serpentine 38 3.5 1.5 to 1 3A 3A Ripped quartz schist (Medium to High) 210 152 69 2.10 0.262 0.6 to 0.9 Good 92 (46 to 142) 97 (94 to 100)
subangular sure if water added to Aug 74)

6p 15t SDVR, 15% 1.44 to >1.8 years (from


Shiroro 125 4.5 1.3 to 1 3A 2 granite (Very High) angular 32 500 22 2.226 0.20 1.0 Good 66 (61 to 71) -
water 5/84)
Tianshengqiao - 1 Limestone, SW to UCS = 70 - 90 15 to 6p 16t SDVR, 20%
178 6.6 1.4 to 1 3A 3B angular 800 - 2.19 0.23 0.8 Good 49 (40 to 57) - -1.5 to >0.8 years
(Zone 3A) FR Very High 20 water
Tianshengqiao - 1 UCS = 16 - 20 6p 16t SDVR, 20%
178 6.6 1.4 to 1 3B 3C Mudstone angular 40 600 20 to 35 2.23 0.21 0.8 Good 37 (32 to 42) - -
(Zone 3B) Medium water
Greywacke, some UCS = 45 angular, 4p 10t SDVR, > 2.05 to 2.35 years (5/81 to
Tullabardine 25 8.6 1.3 to 1 3A 3A 28 400 30.5 2.22 0.23 0.9 to 1.0 Good 74 170
slate High elongated 10% water 8/81)
(High to Very 0.18 to (4p 10t ?) SDVR, > 180 0.18 to 0.24 years (June to
White Spur 43 3.4 1.3 to 1 3A 3A Tuff - SW to FR angular - 1000 - 2.30 1.0 Good 340
High ?) 0.25 10% water (160 to 200) July 1989).
UCS = 66 4-6p 10t SDVR, 1.62 to 5.04 years (6/80 to
Winneke 85 12.4 1.5 to 1 3A 3 SW to FR Siltstone angular 33 800 28 2.07 0.302 0.9 Good 55 (50 to 59) 104
High 15% water 11/83)
UCS = 240 8p 12t SDVR, 25
Xibeikou 95 2.3 1.4 to 1 3A 1 Limestone - FR angular - 600 - 2.18 0.284 0.8 Good 80 (60 to 100) 260 -1.5 to 6 years (June 1995)
Very High to 50% water
(High to Very 4p 10t SDVR, 15% 1.02 to 1.46 (mid to late
Xingo (Zone 3A) 140 6.1 1.4 to 1 3A 3 granite gneiss angular 18 650 4 to 33 2.15 0.28 1.0 Good 34 (30 to 39) 76 (73 to 80)
High ?) water 1994) years
3
Foz Do Areia * 1C & mix basalt & UCS = 235, High 0.8 for 1D 4p 10t SDVR, 25% Good to
160 5.2 1.4 to 1 3B angular 14.2 - - 1.98 0.27 32 (29 to 38) - -
(Zone 3B) 1D basaltic breccia to Very High 1.6 for 1C water Reasonable

Aguamilpa UCS = 180 Good to


185.5 2.6 1.5 to 1 3B 3C ignimbrite angular 22 700 - - - 1.2 4p 10t SDVR 36 (25 to 45) - -
(Zone 3B) Very High reasonable

3 (High to Very (0.33 to 4p 9t SDVR, no


Ita (Zone 3B) * 125 7.0 1.3 to 1 3B E3 Breccia and Basalt angular 13.3 750 15 2.066 1.6 Reasonable 24 (14 to 46) - -
High) 0.39) water
4p 10t SDVR, 15%
Khao Laem UCS < 190
130 7.7 1.4 to 1 3B 3B limestone angular - 1500 - - - 2.0 water lower in 70 Reasonable 30 - -
(Zone 3B) Very High
m
3
Segredo * 1C & basalt (<5% basaltic UCS = 235, High 1D - 0.8 4p 9t SDVR, no 28 (25 to 33),
145 5.0 1.3 to 1 3B angular 10.2 - - 2.01 0.43 Reasonable - -
(Zone 3B) 1D breccia) to Very High 1C - 1.6 water likely arching
Chapter 6: Deformation Behaviour of Rockfill Page 6.31

Table 6.5 (cont.): Summary of embankment and rockfill properties for CFRD case studies (sheet 3 of 3)
Embankment
Material Parameters / Properties of Rockfill
Dimensions Time For First Filling
2
Dam Name 4 Strength* Dry Layer Erf (MPa) (from end of main
Hgt, H Upstream Zone * Rockfill Particle dmax % finer Void Level of Erc (MPa),
L/H 1 Class
n Cu Density Thickness Placement rockfill construction)
(m) Slope Source* Shape (mm) 19 mm 3 Ratio, e Compaction average
Class Auth UCS (MPa) (t/m ) (m)
Sound and
(Medium to Very 6p 10t SDVR, no
Xingo (Zone 3B) 140 6.1 1.4 to 1 3B 4 weathered granite angular 80 750 15 to 60 2.1 0.31 2 Reasonable 13 (12 to 14) - -
High ?) water
gneiss
dumped and well
Courtright 97 2.8 1.15 to 1 3 3 granite (Very High) angular (<7) 1750 - (1.8) 0.47 8 to 52 Poor - - 0 to 9 years (from 1959)
sluiced

El-Infiernillo diorite & silicified UCS = 125 4p D8 dozer, no


148 2.3 1.75 to 1 3B 3B angular 13 600 22 1.85 0.47 0.6 to 1.0 Poor 39 (27 to 48) - -
(Zone 3B) conglomerate Very High water

UCS = 100 - 140, 8 to 2000 to dumped and well 0.07 to 0.58 years (Dec 52
Lower Bear No. 1 75 3.9 1.3 to 1 3 3 granodiorite angular low - - max 65 Poor - 21
very High 10 3000 sluiced to June 53)

UCS = 100 - 140, 8 to 2000 to dumped and well 0.07 to 0.58 years (Dec 52
Lower Bear No. 2 46 5.7 1 to 1 3 3 granodiorite angular low - - max 36.5 Poor - 40
very High 10 3000 sluiced to June 53)
1500 to 8 to 52 dumped and well
Wishon 90 11.3 1.15 to 1 3 3 granite (Very High) angular (<7) - 1.80 0.47 Poor - - 0.4 to 0.45 years (May 1958)
2000 (variable) sluiced

Limestone & Shale dumped and poorly Poor to Very


Dix River 84 3.7 1.1 to 1 3 3 (Very High) angular ? low - - - - 21 - - Started in 1925
(?) sluiced Poor

UCS = 100 - 130, angular, low 2000 to dumped and poorly Poor to Very
Salt Springs 100 4.0 1.3 to 1 3 3 granite low (1.88) 0.41 5 to 52 - 20 0 to 1.48 years (1931/32)
Very High blocky (<10) 3000 sluiced Poor

UCS = 45 7.6 dumped dry (no 2.6 years. Filled in 71 hours


Cogswell 85.3 2.1 1.3 to 1 3 3 Granitic Gneiss angular 7 1300 5 2.05 0.37 Very Poor - 44
High (46 m max) water) (Mar 1938)

El-Infiernillo diorite & silicified UCS = 125 Dumped and


148 2.3 1.75 to 1 3C 3C angular <13 >600 - 1.76 0.54 2.0 to 2.5 Very Poor 22 (17 to 27) - -
(Zone 3C) conglomerate Very High spread, no water

Legend:
1
H = dam height Erc = secant modulus during construction (average)
(average) * FR = fresh, SW = slightly weathered
L = crest length 4p 10t SDVR = 44 passes
passes of 10
10 tonne
tonne smooth drum *2 rock strength classification to AS 1726-1993
Cu = uniformity coefficient (d60/d10) 3
vibrating roller * Brazilian CFRDs constructed with a mix of very high strength basalt, vesicular basalt
dmax = average maximum particle size % water = % by volume and high strength basalt breccia.
4
Erf = deformation modulus during first filling (…) in strength classification, Cu, dry density * Class = For the classification system used (refer Section 6.2.4);
- indicates unknown and void ratio columns indicate estimation Auth = rockfill zoning used by author/s of referenced paper
Chapter 6: Deformation Behaviour of Rockfill Page 6.32

6.4.2.1 Base Plots of Data


Figure 6.15 and Figure 6.16 present the stress-strain and secant modulus-stress
relationships from the internal settlement monitoring records of eleven embankments
(10 CFRDs and El Infiernillo dam). For the ten CFRDs the monitoring records are from
HSGs installed under the embankment centreline and in the lower third to half of the
embankment. For El Infiernillo the monitoring records are from internal vertical
settlement gauges (IVM) in the rockfill shoulders, and include monitoring records
within the dry placed and poorly compacted rockfill zones (Zones 3B and 3C).
Vertical stresses have been estimated from the elastic solutions for embankments by
Poulos et al (1972) and assume that the embankment has been constructed in horizontal
layers across the full embankment width. Use of these solutions takes into
consideration the effect of embankment shape on vertical stress (refer Section 6.2.3).
Note that in Figure 6.15 and Figure 6.16 zero strain is equivalent to the vertical
stress at the mid point of the layer analysed when the upper HSG was placed.
Therefore, in Figure 6.15 the stress-strain curve will not intersect the origin. It will
intersect the vertical stress axis at stresses in the range 30 to 150 kPa. The assumption
of zero strain at a small but finite value of vertical stress introduces a slight error in the
secant moduli calculations.

6.4.2.2 Effect of Compaction/Placement Method

The stress-strain relationships (Figure 6.15) indicate that the degree of compaction has a
significant influence on rockfill modulus, confirming the results of large scale
laboratory tests on rockfill (Section 6.3.2). The dry placed and poorly rockfills of El
Infiernillo dam have lower modulus than for the well-compacted rockfills. The results
also indicate:
• The very high modulus of well-compacted gravels (Crotty dam) in comparison to
well-compacted quarried (or angular) rockfills, although the smaller size and well-
graded particle size distribution of the Crotty gravel fill may also have influenced
the modulus.
• The intact strength (UCS) of the rock used as rockfill has an influence on the
modulus for well-compacted quarried rockfills. The very high strength rockfills
Chapter 6: Deformation Behaviour of Rockfill Page 6.33

(Cethana, Murchison and Reece dams in particular) have greater moduli than the
medium to high strength rockfills used at Tullabardine and Mackintosh dams.

2000

1800
Estimated Vertical Stress (kPa) .

1600

1400

1200

1000

800

600

400

200

0
0.0% 0.5% 1.0% 1.5% 2.0% 2.5% 3.0% 3.5%
Measured Vertical Strain (%)
Cethana Crotty Mackintosh Murchison
Kotmale Reece Bastyan Tullabardine
Khao Laem White Spur El-Infiernillo (loose) El-Infiernillo (dry dumped)

Figure 6.15: Stress-strain relationship for rockfills observed during construction

600
Crotty

500
Secant Modulus (MPa) .

400
Murchison
Khao Laem
300
Cethana
Mackintosh

200

El Infiernillo - loose
100

0
0 200 400 600 800 1000 1200 1400 1600 1800
Estimated Vertical Stress (kPa)
Cethana Crotty Mackintosh Murchison
Kotmale Reece Bastyan Tullabardine
Khao Laem White Spur El-Infiernillo (loose) El-Infiernillo (dry dumped)

Figure 6.16: Secant modulus versus vertical stress from monitoring during construction
Chapter 6: Deformation Behaviour of Rockfill Page 6.34

Particle size distribution and valley shape also influence the stress-strain relationship of
the rockfill shown in Figure 6.15, although they are not readily apparent from the small
number of case studies represented. The effect of these factors is addressed later. Note
that for valley shape, the error in estimation of the vertical stress by ignoring valley
shape effects is then introduced into the stress-strain relationship plotted (i.e. the valley
shape influences the vertical stress in the embankment, and not the vertical stress-strain
relationship of the rockfill).

6.4.2.3 Effect of Confining Stress


An important aspect of rockfill deformation behaviour identified from laboratory testing
is the effect of increasing applied stress. The non-linear stress-strain relationship of
rockfill, indicative of decreasing modulus with increasing applied stress (either
deviatoric, confining or mean normal stress), evident from laboratory oedometer and
triaxial compression testing is clearly evident in the field observations (Figure 6.16 and
Figure 6.17). The initial behaviour of increasing secant modulus at low stress levels
(less than 400 to 600 kPa) for several dams in Figure 6.16 (Murchison and Cethana
dams for example) is considered to be due to seating of the HSG as a result of initial
incompatibility between the filling placed around the instrument and the predominant
rockfill.
Plotting the field data against deviatoric, confining or mean normal stress has not
been attempted due to the introduction of errors in confining stress estimation associated
with estimation of Poisson’s ratio and the observation that Poisson’s ratio is also stress
dependent (Duncan 1996). It is also noted that the stress path represented by the field
observations under the centreline of the embankment closely approximates that of the
oedometer laboratory test (Kovacevic 1994) and the moduli (secant and tangent) would
therefore approximate the confined modulus.
The plot of tangent modulus (Figure 6.17) shows a greater variation in modulus than
for secant modulus plot due to the incremental calculation of tangent modulus.
However, the trend of decreasing modulus with increasing vertical stress is clearly
evident. Over the vertical stress range 500 to 1000 kPa, reductions in the tangent
modulus of approximately 40 to 50% are observed for the well-compacted rockfills.
The variation in modulus (both secant and tangent) for an individual case study is
affected by the accuracy of settlement measurements, the time-dependent or creep
Chapter 6: Deformation Behaviour of Rockfill Page 6.35

deformation of rockfill and therefore the rate of construction, and the variation in
rockfill height across the width and length of the embankment during construction.

6.4.2.4 Effect of Valley Shape

The influence of valley shape is identified (Pinto and Marques Filho 1998; Giudici et al
2000) as a significant factor affecting the vertical stress and therefore the calculated
deformation modulus during construction (from 1 or 2-dimensional analysis),
particularly for embankments constructed in narrow valleys with relatively steep
abutment slopes, because of arching across the abutment slopes. To check these
hypotheses finite difference analysis (FDA), using the FLAC program, was undertaken.

500

400
Murchison
Crotty
Tangent Modulus (MPa) .

300

200
Cethana
Mackintosh
El Infiernillo - loose
100

0
0 200 400 600 800 1000 1200 1400 1600 1800
Estimated Vertical Stress (kPa)
Cethana Crotty Mackintosh Murchison
Kotmale Reece Bastyan Tullabardine
Khao Laem White Spur El-Infiernillo (loose) El-Infiernillo (dry dumped)

Figure 6.17: Tangent modulus versus vertical stress from monitoring during
construction

A two-dimensional FDA was undertaken assuming an idealised rockfill embankment of


100 m height constructed in valleys with river widths of 20, 50 and 100 m, and uniform
abutment slopes of 0, 26.5, 45 and 70 degrees. The rockfill was modelled as a linear
elastic material with Young’s modulus of 100 MPa and Poisson’s ratio (ν) of 0.27, and
the foundation with Young’s modulus of 50 GPa (Figure 6.18). After establishing
Chapter 6: Deformation Behaviour of Rockfill Page 6.36

initial stresses in the foundation, the embankment construction was modelled in 5 m


lifts and stresses (major and minor principal stresses, vertical stress and horizontal
stress) monitored at seven locations (a to g in Figure 6.18) in the centre of the gully.
Grid element sizes of 1.25 m height and 1.25 to 2.5 m width were used. The analysis
was then repeated by construction of the embankment in a single 100 m lift (as was
done in the analysis by Giudici et al. (2000)).
It is recognised that the analysis has several deficiencies, however, for the purpose
of assessment of arching effects on stresses within the embankment the results are
considered reasonably representative. The greatest deficiency in the two-dimensional
analysis is not taking into consideration the significant three-dimensional effect of
embankment shape, which will therefore result in over-estimation of stresses in the
latter stages of construction.

Embankment
35.6 m E = 100 MPa, ν = 0.27
Foundation
E = 50 GPa a
15 m 100 m
ν = 0.27 b Abutment
10 m slope
c
10 m
d
10 m
e
10 m
9.4 m
f
g

River
width

Figure 6.18: Idealised model for two-dimensional finite difference analysis of cross-
valley influence

The results of the analysis are presented in Figure 6.19. Vertical stresses at the end of
construction have been normalised against those for the case with zero abutment slopes
(i.e. no cross-valley arching effect) and plotted against the abutment slope angle. Points
a to g correspond to the locations shown in Figure 6.18.
The results of the analysis indicate:
• Cross-valley arching is only significant (greater than 20% reduction in vertical
stress) for narrow valleys (river width less than 30 to 40% of the dam height) with
Chapter 6: Deformation Behaviour of Rockfill Page 6.37

steep abutment slopes (greater than about 50 degrees, but also dependent on river
width), and then only in the lower third to half of the embankment.
• Where the river width is approximately equal to half the embankment height, cross-
valley arching has some effect (10 to 20% reduction in vertical stress) for abutment
slopes steeper than about 45 degrees, and then only in the lower third to half of the
embankment.
• Negligible arching is observed for river widths equal to about the embankment
height regardless of the abutment slope.
• Modelling embankments in single or very large lifts is likely to result in significant
under-estimation of stresses within the embankment, particularly for river widths
less than about half the embankment height.

(a) 80 River width = 20m, constructed in 5m lifts (b) 80 River width = 20m, constructed in 1 lift

70 70
Abutment Slope Angle (degrees) .
Abutment Slope Angle (degrees) .

Point g
60 60 Point f
Point e
50 50 Point d
Point c
40 40 Point b
Point a
Point g
30 30
Point f
Point e
20 Point d 20
Point c
10 Point b 10
Point a
0 0
0.4 0.5 0.6 0.7 0.8 0.9 1 0.4 0.5 0.6 0.7 0.8 0.9 1
Normalised Vertical Stress Normalised Vertical Stress

(c)
80 River width = 50m, constructed in 5 m lifts (d) 80 River width = 50m, constructed in 1 lift

70 70
Abutment Slope Angle (degrees) .

Abutment Slope Angle (degrees) .

Point g
60 60 Point f
Point e
50 50 Point d
Point c
40 40 Point b
Point g Point a
Point f
30 Point e 30
Point d
20 Point c 20
Point b
10 Point a 10

0 0
0.75 0.8 0.85 0.9 0.95 1 0.75 0.8 0.85 0.9 0.95 1
Normalised Vertical Stress Normalised Vertical Stress
Chapter 6: Deformation Behaviour of Rockfill Page 6.38

(e) 80 River width = 100m, constructed in 5 m lifts (f) 80 River width = 100m, Constructed in 1 lift

70 70
Abutment Slope Angle (degrees) .

Abutment Slope Angle (degrees) .


Point g
60 60 Point f
Point e
50 50 Point d
Point c
Point b
40 40
Point g Point a
Point f
30 Point e 30
Point d
20 Point c 20
Point b
10 Point a 10

0 0
0.9 0.95 1 1.05 0.9 0.95 1 1.05
Normalised Vertical Stress Normalised Vertical Stress

Note: The normalised vertical stress is the vertical stress at the end of construction
normalised against the vertical stress obtained for the flat abutment slope analysis.
Figure 6.19: Results of two-dimensional finite difference analysis of the effect of cross-
valley shape on vertical stresses in the dam. (a), (c) and (e) represent construction in 5
m lifts (to 100 m) and (b), (d) and (f) construction in a single 100 m lift.

Table 6.6 presents an assessment of the effects of cross-valley arching on the


embankments studied as part of this analysis. Significant arching (greater than 20%
reduction in vertical stress) is likely in the lower 25 to 50% portion of the narrow, steep
valley embankment sections of Segredo, Salvajina, Murchison, Cethana and Alto
Anchicaya dams. For Golillas cross valley arching is a significant factor given the
narrow and very steep sided valley in which the embankment has been constructed.

Table 6.6: Assessment of cross-valley influence on arching for case studies analysed
Influence of Cross-Valley Arching
Significant Some Limited to Negligible
(> 20% reduction in Vertical Stress) (10 to 20% reduction in (< 10% reduction in
Vertical Stress) Vertical Stress
Segredo (narrow gully section, lower White Spur (lower quarter) Wishon, Xingo, Xibeikou,
15%) Shiroro (lower third) Winneke, Tullabardine,
Salvajina (lower half in narrow gully) Lower Bear No. 1 (lower Tianshengqiao, Serpentine,
Murchison (lower quarter) quarter) Scotts Peak, Salt Springs,
Golillas Little Para (lower 15%) Reece, Mangrove Creek,
Cogswell (lower quarter) Kangaroo Creek (lower third) Mackintosh, Lower Bear No.2,
Cethana (lower quarter in steep gully) Courtright (lower third) Kotmale, Khao Laem, Ita, Foz
Alto Anchicaya (lower half in steep Bastyan (lower third) do Areia, Dix River, Crotty,
gully) El Infiernillo (lower third) Aguamilpa,
Chapter 6: Deformation Behaviour of Rockfill Page 6.39

The secant modulus versus vertical stress plot during embankme nt construction should
be able to provide an indication of cross-valley arching (for cases where the estimated
vertical stresses have ignored the cross valley influence). The typical plot, assuming no
arching effects, is for a small reduction in the secant modulus with increasing vertical
stress (Figure 6.16). An increase in the calculated secant modulus with increasing
vertical stress is considered to provide an indication of cross-valley arching. The
deformation behaviour for HSG 1 to foundation (lower 10 m or 15%) at Bastyan dam
(Figure 6.20) is indicative of this behaviour. The plot for HSG 1 to 2 (located 50 to 65
m below crest level) is more typical of the normal response. Another indicator could
also be a relatively high calculated secant modulus in the lowest section of the
embankment compared with the section above such as indicated by the monitoring
results at Cethana dam (Figure 6.20). However, this may be affected by variations in
the rockfill properties or placement techniques.

250
Secant Modulus (MPa) .

200

150

100

Cethana (HSG 2)
50 Cethana (HSG 2 to 4)
Bastyan (HSG 1)
Bastyan (HSG 1 to 2)
0
0 200 400 600 800 1000 1200 1400 1600 1800
Estimated Vertical Stress (kPa)

Figure 6.20: Indicators of cross valley arching effects from variations in the secant
modulus with vertical stress.

The analysis shows that valley shape is not as significant as indicated from the empirical
analyses of Pinto and Marques Filho (1998) and Giudici et al (2000), Figure 6.8 and
Figure 6.10 respectively, and the finite element analyses by Giudici et al (2000), Figure
6.11. The differences in calculated deformation moduli for the well-compacted rockfills
must be due to other factors in addition to valley arching, including the rockfill particle
size grading, strength of the rock and type of fill. The analysis also shows that
Chapter 6: Deformation Behaviour of Rockfill Page 6.40

significant errors are introduced into the vertical stress profiles, particularly for
embankments constructed in narrow valleys, by construction of the embankment in a
single stage rather than “building” the embankment in layers.
The method used to calculate the rockfill modulus during construction is also an
important consideration for the empirical analyses. Pinto and Marques Filho (1998) and
Giudici et al (2000) (Figure 6.8 and Figure 6.10) used, as far as the writer is aware, the
simplifying assumption of Figure 6.1a, which would over-estimate the vertical stresses
and therefore over-estimate the calculated moduli.

6.4.2.5 Effect of Particle Size Distribution

The effect of particle size distribution (PSD) on the secant modulus was investigated by
grouping the data into sets with similar degree of compaction, type of fill (rockfill or
gravel) and intact rock strength.
For each rockfill zone (mostly Zone 3A) of each case study the procedure to
estimate a representative secant modulus at the end of construction was:
• Estimation of the vertical stress and secant modulus at the end of construction for
each pairing of HSGs within the rockfill zone allowing for the effect of embankment
shape (refer Section 6.2.3). The secant moduli were, in general, determined from
internal monitoring points in the region bounded by the lower 60% of the
embankment and within the central half of the dam cross section.
• For case studies where valley shape was considered to having had some to
significant influence (Table 6.6) the secant moduli was adjusted by multiplication
with an appropriate stress reduction factor determined from Figure 6.19. Secant
moduli values considered as being greatly affected by valley arching were omitted.
• A representative secant modulus for each rockfill zone was then calculated by
averaging the estimates of secant moduli from that rockfill zone corrected for valley
shape.

The results of laboratory testing on rockfills (Section 6.3.2) indicated that the shape
of the particle size distribution curve (measured by the uniformity coefficient, Cu =
D60/D10) and particle size itself affected the deformation behaviour. To take into
consideration both of these factors various combinations of particle size (e.g. D50, D60,
D80, etc.) and different definitions of uniformity coefficient (D60/D10, D80/D30, etc.) from
Chapter 6: Deformation Behaviour of Rockfill Page 6.41

the average particle size distribution (PSD) for each rockfill zone were statistically
analysed (using single and multiple variable methods) against the representative secant
modulus at end of construction, determined as outlined above. Further statistical
analyses were undertaken incorporating embankment height as a variable in an attempt
to normalise the data against vertical stress.
The analysis indicated that the best predictors of the secant modulus at the end of
construction (Erc) were:
• D80 (diameter for 80% passing from average grading curve) for the very high
strength well compacted quarried rockfill data set (Figure 6.21) with a regression
coefficient of 0.83 for the power function shown.
• D90 (diameter for 90% passing from average grading curve) for the medium to high
strength rockfills with a regression coefficient of 0.70 for the exponential function
given as Equation 6.8. The fit to D80 provided a reasonable correlation with
coefficient of regression of 0.44 (Figure 6.21) for the exponential function shown.
In comparing of the two methods, the difference in calculated moduli in most
instances is less than 5 MPa, or statistically, a standard deviation of less than 7%.
Higher moduli are obtained using the D90 method for PSD with a steep coarse part
of the curve (i.e. above D70) and higher moduli for the D80 method for PSD with a
relatively flat coarse part of the PSD curve typical of gravels. Either approach is
suggested for estimation of Erc for medium to high strength rockfills. For simplicity
D80 has been selected for predictive methods.

E rc = 153 ⋅ e (− 0. 0045⋅ D90 ) (6.8)

where Erc = secant modulus during construction in MPa, D90 = diameter for 90%
passing from average grading curve in millimetres. Equation 6.8 is for medium to high
strength well-compacted rockfills.
From Figure 6.21 it is evident that the secant moduli of the medium to high strength
rockfills shows less variation with increasing D80 than for the very high strength
rockfills. This is considered to reflect the broad, well-graded shape of the particle size
distribution of the medium to high strength rockfills as indicated by the high uniformity
coefficients (Cu values from 45 to 300) in comparison to steeper shape of the very high
strength rockfills (Cu values from 6 to 42).
Chapter 6: Deformation Behaviour of Rockfill Page 6.42

400

Crotty Very high strength, well compacted rockfill


350 Medium to high strength, well compacted rockfill
Gravels, well compacted
Aguamilpa 3A
E rc (MPa), at End of Construction .

300 Reasonable compaction

250

Erc = 113 e(-0.0052 D80)


200 R2 = 0.44
Murchison
150 Erc = 9.0 x 105 . D80 -1.70
Gollilas R2 = 0.83
Cethana
100
Bastyan Reece
50
Salvajina 3B
Scotts Peak Little Para
0
0 50 100 150 200 250 300 350 400 450
D80 in millimetres (particle diameter equivalent to 80 percent passing)

Figure 6.21: Representative secant modulus (mostly Zone 3A rockfill) at end of


construction (Erc) versus D80 from average grading of the rockfill.

Other points to note are:


• The medium strength rockfills of the medium to high strength data set (Little Para,
Scotts Peak and Salvajina Zone 3B) plot below the trendline for the medium to high
strength case studies.
• For the well-compacted gravels there is not enough data to establish a relationship
between D80 and E rc.
• D80, and D90 in the case of medium to high strength rockfills, is a better predictor of
Erc than uniformity coefficient, Cu.
• The predictions are not improved by adding Cu to the equation. This is possibly
related to the partial inter-dependence of Cu and D80, with decreasing Cu generally
observed with increasing D80, particularly for the very high strength rockfills.
• For the very high strength rockfills embankment height (and hence stress level)
appears to have some influence. Apart from Bastyan dam (75 m height) the cases
with embankment heights less than about 120 m plot above the trendline and
embankment heights greater than about 130 m below the trendline. This finding is
not unexpected given the influence of stress level on the secant modulus (Figure
6.16).
Chapter 6: Deformation Behaviour of Rockfill Page 6.43

• No similar correlation with height is evident for the medium to high strength
quarried rockfill data set.

The E rc values in Figure 6.21 for each individual rockfill zone are representative of
the applied vertical stress range of the internal points considered in the estimation of E rc.
The representative mean vertical stress for each data set in Figure 6.21 is given in Table
6.7.

Table 6.7: Stress conditions representative of the data sets from Figure 6.21
No. Vertical Stress Characteristics of Data Points
Data Set Cases Range Mean Comment
(kPa) (kPa)
Very high strength, well- 9 900 to 1400 5 of 9 points with mean
compacted quarried 2200 from 1300 to 1550 kPa
rockfill
Medium to high strength, 9 400 to 800 Even spread of data points
well-compacted quarried 1300 across the stress range
rockfill
Well-compacted gravels 4 950 to 1500 Spread of data points
2200 across the stress range

6.4.2.6 Moduli During Construction of Dumped Rockfill


For dumped rockfill, the limited internal deformation data during construction is
insufficient for analysis and assessment of predictive methods for estimation of the
moduli during construction. The available data records for dumped or poorly
compacted rockfills are:
• At El Infiernillo very high strength diorite and silicified conglomerate were used for
the rockfill. The deformation records indicate Erc values (Figure 6.16) of:
− For the Zone 3B rockfill, placed dry in 0.6 to 1.0 m layers and compacted by
bulldozer, Erc at end of construction ranged from 25 to 49 MPa (average 39
MPa) at vertical stress levels in the range 550 to 1800 kPa.
− For the Zone 3C rockfill, dumped dry and spread in 2 m thick layers, Erc at end
of construction ranged from 17 to 27 MPa (average 22 MPa) at vertical stress
levels in the range 400 to 800 kPa.
Chapter 6: Deformation Behaviour of Rockfill Page 6.44

• At Ita dam (Figure 6.27), the base 10 m thickness of rockfill in the river section was
dumped, presumably through water. E rc at end of construction for the dumped
rockfill (very high strength basalt) was estimated at 15 to 19 MPa at vertical stress
levels in the range 1400 to 2100 kPa.

The tangent moduli for the rockfill at El-Infiernillo (Figure 6.17) approached values
as low as 15 to 20 MPa for the Zone 3B rockfill and 10 to 15 MPa for the Zone 3C
rockfill. Based on back-analysis of the deformations of central core earth and rockfill
dams (Fell et al 2000; Woodward Clyde 1999) the tangent moduli is likely to be lower
for dirty or weathered dumped rockfills, possibly down to as low as 5 MPa at stress
levels exceeding those previously experienced by the rockfill.

6.4.3 DEFORMATION OF THE FACE SLAB OF CFRD ON FIRST FILLING

6.4.3.1 Simplified Methods of Estimation of Maximum Face Slab Deformation

Fitzpatrick et al (1985) and Pinto and Marques Filho (1998) suggest that simple
empirical methods are sufficient to estimate face slab deformations. The most simple
empirical estimates are based on ratios of the modulus on first filling, Erf, to the
deformation modulus during construction, Erc, calculated according to Equation 6.1
(refer Figure 6.1a). Cooke (1993) indicates that the ratio of moduli on first filling to
moduli during construction (Erf / Erc) is typically in the range 1.3 to 3, a range generally
confirmed by most authors discussing the variation.
The calculation of Erf (Fitzpatrick et al 1985) from Equation 6.2 (refer Figure 6.1b)
does not take into consideration the stress path or stress distribution within the upstream
rockfill due to first filling. It is based on the assumption of a uniform stress change,
equivalent to the applied stress from the water load on the concrete face, acting over the
distance normal to the face slab between it and the foundation (Figure 6.1b). Fitzpatrick
et al (1985) recognised that these assumptions were incorrect; but commented that the
method provides a simple and reasonable approximation of the rockfill modulus and
face slab deformation. They indicate that the method is applicable for estimation of Erf
(and therefore prediction of face slab deformation) over the range from 20% to about
60% of the embankment height.
Chapter 6: Deformation Behaviour of Rockfill Page 6.45

Analysis of the data set was undertaken in an attempt to improve on the method of
prediction of the Erf / Erc ratio. Valley shape effects were not specifically taken into
consideration given that valley shape is potentially likely to affect both the secant
moduli at end of construction and modulus during first filling. In the analysis therefore,
the representative secant moduli at end of construction for the Zone 3A rockfill of each
case study are not corrected for valley shape.
The results of the analysis show (Figure 6.22) that there is a reasonable correlation
between Erf / Erc and embankment height taking into consideration the upstream slope
angle. A reasonable estimation of Erf / Erc is possible for upstream slope angles of 1.3 to
1.4H to 1V. For flatter upstream slopes a lower ratio is considered appropriate,
however, there are insufficient data points for estimation of a correlation, although a
very approximate estimation of the trendline is presented in Figure 6.22.

4.0
Khao Laem

3.5 Estimation of trendline for upstream


slopes of 1.3 - 1.4H to 1V

3.0
Scotts Peak
2.5
Erf / E rc

2.0
Mackintosh Ita
Foz do Areia
Golillas
1.5
Crotty 1.3H to 1V
1.0 1.4H to 1V
Serpentine
1.5H to 1V
0.5 Very approximate estimation of trendline for 1.6H to 1V
upstream slopes of 1.5H to 1V 1.7H to 1V
0.0
0 25 50 75 100 125 150 175 200
Embankment Height (m)

Figure 6.22: Erf / Erc ratio versus embankment height

A number of cases plot away from the trendline for the 1.3 –1.4H to 1V data set or are
not in accordance with the general trend of decreasing Erf / Erc ratio with flattening of
the upstream slope. The reasons for these observations are:
• For Mackintosh and Foz do Areia the face slab deflection profiles are different to
the typical profile (refer Section 6.4.3.2) with maximum deflection measured at 15
to 20% of the dam height above the toe; i.e. lower than normal.
Chapter 6: Deformation Behaviour of Rockfill Page 6.46

• For Ita dam the dumped rockfill in the river section (Figure 6.27), which is of low
secant modulus in comparison to the Zone 3A compacted rockfill, is considered to
have influenced the face slab deflection.
• At Crotty, the facing zone between the main gravel rockfill and the concrete face
comprised quarried rockfill. The likely lower modulus of this facing zone is
considered to have had an effect of the deformation behaviour of the face slab
resulting in the relatively low value of the Erf / Erc ratio.
• At Scotts Peak the embankment zoning (Figure 6.28) is considered to have affected
the calculation of Erf, which was based on the measured settlement in the HSGs
adjacent to the upstream face slab.
• Serpentine dam plots at quite a low ratio. The reason for this is uncertain.
• At Khao Laem the high value of Erf, and hence the high Erf / Erc ratio, was
considered by Watakeekul et al (1985) to possibly be due to arching effects at the
maximum section within a localised deep zone confined to upstream of the
embankment centreline. Therefore, arching effects influenced Erf and not the secant
modulus at the end of construction, Erc (i.e. they influenced one but not both moduli,
which is unusual).

Other factors considered to affect the Erf / Erc ratio are:


• Fitzpatrick et al (1985) considered the time to completion of first filling a factor, the
shorter the time of first filling the greater the likely ratio due to time dependent
strain effects.
• Cooke (1984) suggested that the compacted rockfill might be substantially stiffer in
the horizontal direction, due to the greater compaction of the upper compared to the
lower sections of a compacted layer. A two-dimensional finite difference analysis,
using the computer program FLAC, was undertaken on Reece dam to further
evaluate the effect of intra-layer differential moduli (differential in the both vertical
and horizontal directions) on the face slab deformation behaviour. The embankment
was modelled in 1 m layers with a 3 to 1 ratio in moduli between the upper 0.5 m
and lower 0.5 m of each layer calculated such that the overall vertical modulus of
any layer was equivalent to that measured from actual observations during
construction. The modelling indicated that the deforma tions on first filling were
similar, less than 5% variation, with or without consideration of variable inter-layer
modulus. In addition, finite element analysis by Kovacevic (1994), using and an
Chapter 6: Deformation Behaviour of Rockfill Page 6.47

elasto-plastic constitutive model (modelling pre-peak plasticity), obtained


reasonable predictions of face slab deformation on first filling by for Winscar and
Roadford dams (refer Section 6.3.3.1). It is concluded, therefore, that the effect of
intra-layer differential moduli has a negligible effect on the face slab deformation
behaviour during first filling.
• Valley shape may influence the Erf / Erc ratio to a small extent. For the 1.3H to 1V
cases, the CFRDs constructed in narrow valley (L/H < 3.2, the ratio of crest length,
L, to the embankment height, H) tend to plot above the trendline whilst for the
broader valleys tend to plot below the trendline, although there are exceptions. For
the 1.4H to 1V cases the trend with respect to valley shape is not as evident.

Two-dimensional finite difference analysis was undertaken to evaluate the stress


paths of the rockfill in the upstream zone of the embankment for different upstream
slopes. Analyses were undertaken on a nominal 100 m high embankment with upstream
slopes of 1.3H to 1V and 1.55H to 1V. The embankment was constructed in 5 m lifts
and the reservoir impoundment in 6 stages to a maximum height of 96 m (i.e. assuming
4 m of freeboard). Given that the interest was in the stresses within the embankment,
the rockfill was modelled as a linear elastic material with Young’s modulus of 100 MPa
and Poisson’s ratio of 0.27. The concrete face slab was not modelled and the upstream
facing zone was modelled with similar modulus parameters to the main rockfill. It is
recognised that the assumption of linear elastic properties would lead to some errors in
the stress distribution, however, the purpose of the analysis was for comparison and
errors were likely to be relatively minimal.
Stresses were calculated at 16 locations in the upstream zone and under the
embankment centreline (Figure 6.23 and Figure 6.24). The stress paths for the points
normal to the face slab at a height 30% above the embankment toe are presented in
Figure 6.23 for the 1.3H to 1V upstream slope angle case and Figure 6.24 for the 1.55H
to 1V upstream slope angle case. The results indicate that:
• The stress paths during construction (Figure 6.23b and Figure 6.24b) show a similar
trend, with some variation, for both the 1.3H to 1V and 1.55H to 1V upstream slope
angle. The general trend of the stress path up to the end of construction is for
increasing deviatoric and mean normal stress.
• In the early stages of first filling the deviatoric stress decreases whilst the mean
normal stress continues to increase. In the mid to latter stages of filling the
Chapter 6: Deformation Behaviour of Rockfill Page 6.48

deviatoric stress starts to increase again reaching stress levels in excess of those at
the end of construction.
• The overall increase in deviator stress from the end of construction to the end of first
filling is greater (at all comparison points in the upstream shoulder) for the 1.55H to
1V upstream slope case, whilst the overall increases in mean normal stress are
relatively uniform.
• The difference in stress path from the end of construction to the end of first filling is
likely to result in greater strains in the 1.55H to 1V case on first filling and therefore
a reduced deformation modulus on first filling (Erf). Given that the secant modulus
at the end of construction (Erc) will be similar between the cases, the ratio Erf / Erc is
likely to be smaller for the 1.55H to 1V upstream slope. Estimation of Erf from the
deformation of the upstream face during first filling modelling confirms this.

The finite difference analysis confirms the findings from field observations (Figure
6.22) that the upstream slope angle is likely to affect the Erf / Erc ratio. Flatter upstream
slope angles will result in a reduced Erf / Erc ratio all other factors being equal.
It is important to recognise that Erf is not a proper simulation of the rockfill modulus
during first filling rather it is an artefact of the method of calculation (refer Figure 6.1b).
This method uses simplifying assumptions in regard to the geometry of the section
analysed, and the assumed stress increase due to the water is not properly modelling the
stresses, as seen in Figure 6.23b and Figure 6.24b. Erf should only be used to calculate
face slab deflections.
Chapter 6: Deformation Behaviour of Rockfill Page 6.49

(b) 500

Principal Deviatoric Stress, q (kPa) .


Points normal to face slab at
30% of dam height
400

300

200 Point f
Point g
Point h
Point i
100
Point o
end of construction
0
0 200 400 600 800 1000 1200
Mean Normal Stress, p (kPa)
Figure 6.23: Stress paths during construction and first filling for nominal 100 m
embankment with 1.3H to 1V upstream slope angle; (a) monitoring point locations, (b)
stress paths for points normal to face slab at 30% of the embankment height.

500
(b)
Principal Deviatoric Stress, q (kPa) .

Points normal to face slab


at 30% dam height
400

300

Point f
200 Point g
Point h
Point i
100 Point o
end of construction

0
0 200 400 600 800 1000 1200
Mean Normal Stress, p (kPa)
Figure 6.24: Stress paths during construction and first filling for nominal 100 m
embankment with 1.55H to 1V upstream slope angle; (a) monitoring point locations, (b)
stress paths for points normal to face slab at 30% of the embankment height.
Chapter 6: Deformation Behaviour of Rockfill Page 6.50

6.4.3.2 Shape of the Face Slab Displacement


The deformed shape of the concrete face on first filling is strongly dependent on the
zoning of rockfill in the body of the embankment. For embankments constructed almost
entirely of a single rockfill zone (Figure 6.25) the typical pattern of deformation is for
maximum deflection at about 30 to 50% of the embankment height reducing to zero at
the toe and with constant to gradual reduction in deformation toward the crest.

Figure 6.25: Face slab deflection during first filling (4/2/71 to 25/4/71) of Cethana dam
(Fitzpatrick et al 1973).

Where the embankment zoning comprises rockfill zones of differing modulus the
deflected shape of the face slab can reflect the modulus variations. In cases where large
differences in rockfill modulus occur, such as the use of gravels of very high modulus
with quarried rockfills of significantly lower modulus, problems with cracking of the
face slab due to development of tensile stresses can occur resulting in leakage. Several
examples of the effect of rockfill modulus variation on face slab deflection are:
• Aguamilpa dam (Mori 1999), Figure 6.26. The face slab deflection (Figure 6.26)
was significantly affected by the modulus variation between the upstream gravel
zone and downstream zone of quarried rockfill. The maximum deflection was
measured near to the embankment crest. And the resultant deformation profile was
considered by Mori (1999) to have caused cracking in the upper section of the face
slab.
• The typical design detail for several large Brazilian CFRDs (Ita dam Figure 6.27)
incorporates the better quality quarried rockfill placed in relatively thin layers in the
upstream third of the embankment and the poorer quality rockfill placed in larger
lifts in the central and downstream regions resulting in a modulus variation. Ita dam
incorporated a zone of dumped rockfill in the river section resulting in larger
deflections where the zone of influence incorporated these weaker rockfill zones.
Chapter 6: Deformation Behaviour of Rockfill Page 6.51

• A similar pattern of face slab deflection was observed at Mangrove Creek dam due
to modulus variation between the fresh sedimentary rockfill used close to the
upstream face and the main body of weathered sedimentary rockfill.
• At Scotts Peak dam (Figure 6.28) the use of rockfill zones with significantly
different moduli contributed to cracking of the asphaltic concrete membrane face.
The embankment design incorporated a zone of well-compacted gravels at the
upstream toe and main body of argillite rockfill of medium to high intact rock
strength. On first filling, large differential deflections caused tensile cracking of the
upstream face near to the contact between the gravel and argillite rockfill zones
resulting in leakage flows in the order of 100 litres/sec.

Figure 6.26: Face slab deformation during first filling of Aguamilpa dam (Mori 1999)

Figure 6.27: Face slab deformation during first filling of Ita dam (Sobrinho et al 2000)
Chapter 6: Deformation Behaviour of Rockfill Page 6.52

Figure 6.28: Scotts Peak dam; embankment zoning and location of face cracks (courtesy
of Hydro Tasmania)

Figure 6.29: Mackintosh dam (a) embankment section (courtesy of Hydro Tasmania)
and (b) face slab deformation on first filling (Knoop and Lack 1985)
Chapter 6: Deformation Behaviour of Rockfill Page 6.53

Face slab deflection profiles different to the typical profile (Figure 6.25) were also
observed at Mackintosh (Figure 6.29) and Foz do Areia dams. In both cases the
maximum deflection during first filling was measured at 15 to 20% of the dam height
above the toe, a much lower level than typically observed. The reasons for this
observation are not clear, however, it is suspected that it is likely to be related to the
rockfill modulus. Potentially lower modulus zones may have been present in lower
portion of these embankments due to the initial production of rockfill of weaker intact
rock strength or poorer gradation (low Cu or larger particle size), or placement at a
lower relative density.

6.4.3.3 Prediction of Face Slab Deflection on First Filling


For the more typical CFRD geometries (Figure 6.25) or where large variations in
rockfill moduli are not likely between rockfill zones, simple empirical correlations are
considered sufficient for prediction of face slab deflection on first filling, using Erf
values estimated from Figure 6.22. Details on the suggested procedure for empirical
estimation of deformations are discussed in Section 6.5.2.
Mori (1999) considers that a more rigorous method of determination of the face slab
deformation than the method by Pinto and Marques Filho (1998) is required. He states
that it should take into consideration the geometry of the structure, the stress path on
impounding and the strength and deformation of the various rockfill zones within the
embankment.
For prediction of face slab deformation behaviour during first filling for complex
rockfill zoning geometries and/or with large moduli variations between the rockfill
zones, finite element analyses should be considered. The detailed analysis would be
justified considering the potential for face slab cracking as a result of possible
development of tensile stresses.

6.4.4 POST CONSTRUCTION CREST SETTLEMENT

Available methods for prediction of post-construction crest settlement of rockfill


embankments (see Section 6.3.3.2) are empirical and are generally based on historical
records of similar embankment types and similar methods of construction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.54

An important aspect of the post construction deformation behaviour of rockfill


related to its stress-strain characteristics is that relatively large deformations occur on
application of stresses above those not previously experienced by the rockfill, such as
on first filling or embankment raising. On un-loading or re-loading to stress states less
than those previously experienced (such as due to fluctuations in the reservoir level) the
rockfill moduli is very high and the resultant deformation therefore likely to be minimal.
Collapse deformation on initial wetting can also result in relatively large deformations.
The following discussion will consider post-construction deformation behaviour of
rockfill (predominantly crest settlements) in CFRD due to:
• Events where stresses are likely to exceed those previously experienced, most
notably first filling;
• Ongoing, time-dependent (or creep) deformation of rockfill; and
• Deformations due to wetting from rainfall or leakage.

Two significant factors in using empirical methods for prediction of the crest
settlement post construction are the base time of deformation, and consideration of the
timing of events such as first filling. If near full impoundment occurs before
completion of construction or before monitoring points are established, then a
significant component of the deformation is likely to have occurred and will not be
captured by the post-construction monitoring. It is important that this be considered in
comparison between case studies and the forward prediction of future deformations. It
is also important that the base time of deformation is clearly defined in the predictive
method being used and that the same definition is used when forward predicting future
deformations.

6.4.4.1 Case Study Records of Post-Construction Crest Settlement

The typical post-construction crest settlement pattern, ignoring first filling effects, is for
crest settlement to occur at a decreasing rate with time. Most equation based empirical
methods for prediction of post-construction deformation model this behaviour by
logarithmic relationships of strain versus log time (Sowers et al 1965) or power type
relationships of strain (Soydemir and Kjaernsli 1979) or strain rate (Parkin 1977) versus
time.
Chapter 6: Deformation Behaviour of Rockfill Page 6.55

Total post-construction crest settlements will be dependent on when the base


reading of mo nitoring points is established after the end of construction. Differences in
terms of even months to establish base readings after the end of construction are
significant as strain rates are greatest immediately after construction.

6.4.4.1.1 Zero Time, to, for Post Construction Crest Settlement

The method proposed by Parkin (1971) (refer Figure 6.12 in Section 6.3.3.3) was used
for evaluation of the initial or zero time, to, for start of post-construction crest
deformation for each case study. The basis of Parkin’s method, derived from rate
process theory, is the assumption that strain rate versus time is a power law relationship
and therefore a plot of log strain rate versus log time (Figure 6.12b) will be linear
provided to is correctly established.
Figure 6.30 shows examples of the time versus inverse strain rate plots for four
dams, showing that for three cases to at the end of the main rockfill construction is the
best fit of the data, and in one case (Crotty dam) to at the end of first filling the best fit.
It was found that the best fit to each case study was for a to at either the end of the main
rockfill construction or at close to the completion of first filling. In summary, the
analysis indicated that:
• For CFRD constructed of dumped rockfill (excluding Salt Springs dam) and for
well-compacted CFRD generally less than 100 m in height (excluding Crotty dam)
the best fit for to was at the end of construction.
• For well-compacted CFRD generally greater than 100 m in height as well as Salt
Springs dam (dumped rockfill CFRD of 100 m height) and Crotty dam (83 m high
well-compacted CFRD constructed of river gravels), the best fit for to was at the end
of first filling.

The end of main rockfill construction is defined as the time of completion of the
main rockfill zones (Zones 3A and 3B). In some cases this may be a significant time
(up to 1 to 2 years) before actual completion of the concrete face and crest detail works.
Chapter 6: Deformation Behaviour of Rockfill Page 6.56

20

Time (years since end of


16

construction)
12

Crotty
4 Murchison
Little Para
Reece
0
0 100 200 300 400 500 600 700
1/(Strain Rate), (year per % vertical crest strain)

Figure 6.30: Examples of derivation of zero time for post-construction settlement

For consistency it is considered necessary for the initial or zero time, to, to be at either
the end of main rockfill construction or at the completion of first filling. For zero time
at the end of main rockfill construction (Figure 6.31 to Figure 6.33), the power function
of the slope of the log strain rate versus time log plot (Equation 6.5 and Figure 6.12b)
was closer to 1, and therefore the long-term slope of the vertical crest strain versus log
time was close to linear (with some exceptions) allowing for easier prediction of future
deformation. However, visual interpretation for comparative purposes is difficult due of
the variation in time between end of construction and start of monitoring for the cases
and the deformations during first filling.
For zero time, to, at the end of first filling (Figure 6.34 to Figure 6.36) visually the
comparison between case studies is much clearer and upper and lower bounds for a
specific set of historical monitoring records more readily identified, allowing for easier
evaluation of “abnormal” deformation behaviour. However, future predictions are
visually more difficult given the curved nature of the deformation behaviour in the plots
of crest settlement versus log time. This curvature is due to the power function, m, in
Equation 6.5 being less than 1 for most case studies.

6.4.4.1.2 Shape of the Vertical Crest Settlement Plots


For the vertical crest settlement versus log time plots for zero time at the end of the
main rockfill construction (Figure 6.31 to Figure 6.33) two features are readily
observed; the ‘S’ shape behaviour for a number of the case studies, and the effect of the
Chapter 6: Deformation Behaviour of Rockfill Page 6.57

time difference between the end of rockfill construction and the initial crest settlement
reading. The ‘S’ shaped curve is a feature of those cases where first filling was
commenced more than about 0.3 to 0.5 years after the end of the main rockfill
construction and the embankment height was greater than about 50 m. The steep
portions of the curve are associated with crest settlement during first filling due to the
increase in stress levels in the body of the rockfill beyond levels previously experienced.
Steeper sections of the slope are generally observed for those case studies where the rate
of filling was relatively rapid and where first filling started at least several years after
construction. At the end of first filling the strain rate remains at a relatively high rate
(on log time scale) due to time dependent effects and then decreases to a strain rate
typical of the long-term deformation rate (on a log scale).

0.0%
Crest Settlement (% of dam height) .

0.2%
Acceleration due to raising of
0.4% full supply level by 3.05 m

0.6%

0.8%
Wishon
Dix River
1.0%
Lower Bear No. 1
1.2% Lower Bear No. 2
Salt Springs
1.4% Courtright
end of first filling
1.6%
0.1 1.0 10.0 100.0
Time (years since end of main rockfill construction)
Figure 6.31: Post-construction crest settlement versus time for dumped rockfill CFRD,
to at end of main rockfill construction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.58

a) 0.0%

Crest Settlement (% of dam height) .


0.2%

0.4%

0.6%
Scotts Peak (ff=0.42 to 2.6)
Mackintosh (ff=1.9 to 2.9)
0.8% Mangrove Creek (ff > 15)
Xingo (ff=1.02 to 1.46)
Tianshengqiao (ff > 0.8)
1.0%
End of first filling
Note: ff = first filling
1.2%
0.1 1.0 10.0 100.0
Time (Years since end of main rockfill construction)

b) 0.00%
Crest Settlement (% of dam height) .

0.05%

0.10%

0.15%

Winneke (ff=1.6 to 5.4)


0.20%
Kangaroo Creek (ff=0 to 1.95)
Serpentine (ff=0.24 to 2.9)
0.25% Tullabardine (ff=2.0 to 2.35)
Little Para (ff=0.6 to 2.8)
White Spur (ff=0.18 to 0.24)
0.30%
end of first filling
Note: ff = first filling
0.35%
0.1 1.0 10.0 100.0
Time (years since end of main rockfill construction)

Figure 6.32: (a) and (b) Post-construction crest settlement versus time for CFRD
constructed of compacted rockfills of medium to high intact strength, to at end of main
rockfill construction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.59

a) 0.00%
Foz Do Areia (ff=0.5 to 0.9)
Crest Settlement (% of dam height) . Reece (ff=1.3 to 1.46)
0.05% Shiroro (ff > 1.8)
Kotmale (ff=0.54 to 1.04)

0.10% Aguamilpa (ff=0.48 to 1.83)


end of first filling

0.15%

0.20%

0.25%

Note: ff = first filling


0.30%
0.1 1.0 10.0 100.0
Time (years since end of main rockfill construction)

b) 0.00%
Crest Settlement (% of dam height) .

0.02%

0.04%

0.06%

0.08%
Alto Anchicaya (ff=0.18 to 0.2)
Bastyan (ff=0.86 to 1.05)
0.10%
Cethana (ff=0.26 to 0.48)
Murchison (ff=1.04 to 1.09)
0.12% Golillas (ff=4 to 5.2)
Crotty (ff=0.87 to 2.0)
end of first filling Note: ff = first filling
0.14%
0.1 1.0 10.0 100.0
Time (years since end of main rockfill construction)

Figure 6.33: (a) and (b) Post-construction crest settlement versus time for CFRDs
constructed of well-compacted rockfills of very high intact strength and of well-
compacted gravels, to at end of main rockfill construction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.60

0.0%
Acceleration due to raising of
Crest Settlement (% of dam height) . 0.2%
full supply level by 3.05 m

0.4%

0.6%

Wishon
0.8% Dix River
Lower Bear No. 1
1.0% Lower Bear No. 2
Salt Springs
1.2% Courtright

1.4%
0.1 1.0 10.0 100.0
Time (years since end of first filling)

Figure 6.34: Post-construction crest settlement versus time for dumped rockfill CFRD,
to at end of first filling.

0.00%
Crest Settlement (% of dam height) .

0.05%

0.10%

0.15%

0.20%

0.25% Winneke
Kangaroo Creek
0.30% Scotts Peak
Serpentine
0.35% Tullabardine
Little Para
0.40% Mackintosh
Mangrove Creek
0.45% Xingo
White Spur
0.50%
0.1 1.0 10.0 100.0
Time (years since end of first filling)
Figure 6.35: Post-construction crest settlement versus time for CFRD constructed of
compacted rockfills of medium to high intact strength, to at end of first filling.
Chapter 6: Deformation Behaviour of Rockfill Page 6.61

0.00%
Alto Anchicaya
Bastyan
0.02% Cethana
Crest Settlement (% of dam height) .
Foz Do Areia
Murchison
0.04% Reece
Kotmale
Golillas
0.06% Crotty
Aguamilpa

0.08%

0.10%

0.12%

0.14%
0.1 1.0 10.0 100.0
Time (years since end of first filling)

Figure 6.36: Post-construction crest settlement versus time for CFRDs constructed of
well-compacted rockfills of very high intact strength and of well-compacted gravels, to
at end of first filling.

From analysis of the deformation during first filling it was observed that the amount of
crest settlement during the period of first filling was dependent on a number of factors
(Table 6.8). The settlement (as a percentage of embankment height) was found to
increase with/for:
• Increasing time of first filling;
• Increasing embankment height;
• Dumped compared to compacted rockfills;
• Decreasing intact rockfill strength for compacted quarried rockfills. For
embankments constructed predominantly of gravels the crest settlements (as a
percentage of embankment height) were similar to those of the very high strength
rockfills; and
• The closer the start of first filling to the end of the main rockfill construction. This
was considered to be due to the concurrent process of time dependent deformation,
the rate for which decreases with time after the end of construction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.62

Table 6.8: Summary of crest settlement during the period of first filling
Rockfill Placement No. of Time of First Embankment Range of Crest
Method and Intact Cases Filling (years) Height Range Settlement
Rock Strength (m) (% of embankment
height)
Dumped rockfill 6 Up to 1 year 43 to 100 0.15 to 0.48
> 1 year 0.23 to 0.34
Compacted gravels 2 1.13 to 1.2 83 and 125 0.017 to 0.024
< 0.1 yrs 43 to 140 0.010 to 0.014
Compacted very high 0.1 to < 0.5 year 75 to 110 m 0.020 to 0.027
10
strength rockfill > 110 m 0.053 to 0.20
> 0.5 year 90 and 166 0.030 to 0.13
< 0.5 year 25 0.008
Compacted medium to
7 > 1 year 38 to 53 0.058 to 0.096
high strength rockfill
60 to 85 0.058 to 0.129

The time between the end of rockfill construction and the initial crest settlement reading
has a significant influence on the measured crest settlement. Relatively large magnitude
total crest settlements (as a percentage of embankment height)are observed when initial
readings are taken shortly after the end of construction (Tianshengqiao in Figure 6.32a
and Kotmale in Figure 6.33a). Where the time delay to monitoring is relatively large
(more than about four to six months after the end of rockfill construction), the
magnitude of total crest settlement is smaller than for similar type CFRD (e.g. Gollilas
in Figure 6.33b and Mangrove Creek in Figure 6.32a). As previously discussed, this
effect makes for difficulties when attempting to compare the deformation behaviour of
CFRD.
These effects are not as evident for the crest settlement plots for zero time at the end
of first filling (Figure 6.34 to Figure 6.36) for which upper and lower bounds of
deformation are clearer and can be more meaningfully defined.
The values of the coefficient, m, in the strain rate versus time power function
(Equation 6.5) was investigated for both to at the end of the main rockfill construction
and at the end of first filling. Results are given in Table 6.9. The value of m represents
the slope of the log strain rate versus log time linear relationship.
The findings indicated some dependence of m on the method of placement and
intact rockfill strength with lower values for the poorly sluiced dumped rockfill.
Chapter 6: Deformation Behaviour of Rockfill Page 6.63

Overall the degree of variability within any data set was quite large and no readily
identifiable correlation was apparent for prediction of m.

Table 6.9: Values of the coefficient, m, in the strain rate – time power function
(Equation 6.5).
Rockfill Placement Power Coefficient, m Comments
Method and Intact to at end of to at end of
Strength rockfill first filling
construction
Dumped and poorly 0.61 and 0.88 0.50 and Limited number of cases (6), only
two poorly sluiced.
sluiced 0.57
Dumped and well sluiced 1.21 to 1.84 0.78 to
0.97
Compacted medium to 0.49 to 1.44 0.29 to More consistent values for to at
end of first filling; 6 of 7 cases in
high strength quarried 0.69
the range 0.42 to 0.69. Greater
rockfill spread for to at end of rockfill
construction.
Compacted very high 0.92 to 1.84 0.62 to Broad spread of values for to at
end of rockfill construction.
strength quarried rockfill 1.04
Better consistency for to at end of
first filling; 7 of 8 cases in the
range 0.82 to 1.09.
Compacted gravels 1.77 and 1.86 0.96 (both) Only two cases, but reasonably
consistent values.

6.4.4.1.3 Summary of the Factors Affecting the Post Construction Crest


Settlement Behaviour

In summary, the most dominant factors influencing the post-construction crest


settlement (as a percentage of the embankment height) are:
• The method of placement (dumped versus compacted). Significantly greater crest
settlements are evident for the dumped and sluiced rockfills compared with
compacted rockfills. The rates of crest settlement (on a log scale) post first filling
are also significantly greater for the dumped rockfills.
• Intact rockfill strength for compacted rockfills. For embankments constructed of
medium to high intact strength rockfill, the total magnitude of settlement at 10 years
is on average approximately twice that of very high strength rockfills. The long-
term, post first filling rates of crest settlement (on a log scale) are also significantly
Chapter 6: Deformation Behaviour of Rockfill Page 6.64

higher for the medium to high strength rockfills, 2 to 10 times that of the very high
strength rockfills.
• Use of gravels as opposed to rockfill. For embankments constructed of
predomi nantly well-compacted gravels (Crotty and Golillas) the total post-
construction settlement is less than that for the well-compacted, very high strength
rockfills. This is likely to be due to several reasons, but is considered to be mainly a
result of the rounded shape of the gravels. The point area of contact between
particles of rounded gravel will be significantly greater than that of angular quarried
rockfill. Hence, the contact stresses will be significantly less resulting is less
particle breakage.
• Embankment height. The long-term rate of crest settlement (on a log scale) is lower
for the smaller embankments within each data set; e.g. Lower Bear No. 2 of the
dumped rockfill cases; Tullarbardine, Serpentine and Little Para dams of the
compacted medium to high strength rockfill cases; and Bastyan and Murchison
dams of the compacted very high strength rockfill cases. The effect of embankment
height is probably not as significant for the compacted very high strength rockfill
cases.

Other factors that appear to affect the post construction crest settlement are:
• The poorly sluiced dumped rockfills (Salt Springs and Dix River dams) tend to have
greater rates of long-term crest settlement (on a log scale) than for well-sluiced
dumped rockfills. The values of the power coefficient m are also lower for the
poorly sluiced dumped rockfills.
• Particle size and uniformity coefficient are thought to have some influence on post-
construction crest settlement behaviour. Greater long-term crest settlement rates (on
a log scale) are likely for dams with coarser rockfills. However, no statistical
correlation is evident from the data indicating that its influence is relatively minor in
relation to other factors.
• Embankment zoning geometry of the main rockfill. Given that the Zone 3B rockfill
is generally of lesser quality compaction (and sometimes lesser quality rockfill
materials) for typical CFRD designs, greater crest settlement rates (per log cycle of
time) would be expected for increasing size of the Zone 3B.
Chapter 6: Deformation Behaviour of Rockfill Page 6.65

• Amount of fluctuation in reservoir level. It is suspected, although the data does not
indicate, that the greater cyclic stress changes associated with larger fluctuations in
reservoir level could result in greater overall settlements.
• Climate. In higher rainfall areas greater amounts of water infiltration into the
rockfill would be expected, potentially resulting in greater settlements.

6.4.4.1.4 Prediction of Post Construction Crest Settlement


Prediction of the post-construction crest settlement is relatively difficult given the
variable influence of first filling and the significant effect of timing on the initial base
reading. However, reasonable predictions are possible.
The proposed method is to consider separately the time dependent deformation and
the settlement due to first filling, as shown in Figure 6.37 for Bastyan dam for example.
During (and shortly after) the period of first filling, the time dependent deformations are
assumed to occur concurrently with the crest settlements due to increase in vertical
stresses from the water load. Estimation of the settlement at some time after the end of
main rockfill construction is therefore by summation of the time dependent and first
filling related components.

0.00%
Start of first filling
Crest settlement (% dam height) .

0.01%

0.02% Effect of first End of first filling


filling
0.03%

0.04%

0.05%

Time dependent
0.06%
deformation
0.07%
0.1 1.0 10.0
Time (years since end of main rockfill construction)

Figure 6.37: Post construction crest settlement of Bastyan dam

(a) Time dependent crest settlement


Figure 6.38 presents the data for long-term post first filling crest settlement rate (percent
per log cycle of time) for the compacted rockfill embankments, for to at the end of main
Chapter 6: Deformation Behaviour of Rockfill Page 6.66

rockfill construction. Trendlines with reasonable to good regression coefficients have


been derived for compacted high strength rockfills and for compacted very high strength
rockfills and gravels. The intact strength classification is that of the dominant rockfill
zone under the dam axis, which in most cases is the Zone 3A rockfill, but in several
cases is the Zone 3B rockfill. A tentative trendline is also given for the medium
strength rockfills.
Kotmale and Mangrove Creek dams stand out as outliers to the general trend. For
Mangrove Creek dam this is thought to be due to the fact that first filling to full supply
level has yet to be reached after 15 years, and for which the long-term settlement rate
was estimated from the settlement data more than 3 years after the end of construction
(refer Figure 6.32a). For Kotmale dam the reason is not clear, although the available
crest settlement data is limited to about 1 year after first filling, and it is possible that the
rate (per log cycle of time) may have since reduced.

0.30

Kotmale
0.25
Settlement (%/log cycle) .
Long-term Rate of Crest .

Mangrove Creek
2
R = 0.79
0.20

0.15

0.10

R2 = 0.51
0.05

0.00
0 20 40 60 80 100 120 140 160 180
Embankment Height (m)
Gravels Very high strength rockfills
High strength rockfills Medium strength rockfills

Figure 6.38: Long-term crest settlement rate (as a percentage of embankment height per
log cycle of time) versus embankment height for compacted rockfills

For the dumped rockfill CFRD the data (Figure 6.31) indicates an increasing rate of
vertical crest strain (per log cycle of time) with time. Estimates of the long-term
vertical crest strain (per log cycle of time) can be derived from either Figure 6.31 or
Table 6.10 based on the time period after the end of main rockfill construction.
Chapter 6: Deformation Behaviour of Rockfill Page 6.67

Table 6.10: Estimates of long-term crest settlement rates for dumped rockfills
Post First Filling Crest Settlement Rate
Time Period (from
(settlement as a % of embankment height per
end of main rockfill
log cycle of time)
construction)
Range Mean
0.5 to 5 years 0.10 to 0.58 0.27
5 to 20 years 0.25 to 1.14 0.66
20 years plus 0.33 to 1.44 0.85

(b) Crest settlement under increased stresses from first filling


Whilst a reasonable estimate of the crest settlement rate per log cycle of time (as a
percentage of embankment height) can be made for the time dependent component of
crest settlement from Figure 6.38 and Table 6.10, estimates of crest settlement during
first filling are more difficult. Two alternatives methods are proposed:
(i) Estimation of the crest settlement attributable to first filling which covers the
period of first filling and the short time thereafter as shown in Figure 6.37. Figure
6.39 summarises the crest settlement (as a percentage of embankment height)
attributable to first filling and can be used for predictive purposes based on the
intact strength of the rockfill, placement method (either dumped or compacted)
and embankment height. The relatively broad variation for some groups, in
particular for dumped rockfill, reflects the high degree of variability from the case
studies (refer Section 6.4.4.1.2). Trendlines are provided for the compacted
gravels and rockfill.
(ii) Estimation of the crest settlement during first filling from Table 6.8. The
estimates from Table 6.8 are different to those from Figure 6.39 in that they are
inclusive of the time dependent settlement component and only cover the period
during first filling (i.e. they do not consider the higher rate of crest settlement
observed that occurs for some CFRDs after the end of first filling). When using
Table 6.8 the time dependent settlement component during the period of first
filling should not be added to that from the Table 6.8.

There is a high degree of variability in using either method, particularly for the
dumped rockfill case studies. Interpretation of crest settlement attributable to first
filling in Figure 6.39 (from Figure 6.31 to Figure 6.33) was imprecise for a number of
Chapter 6: Deformation Behaviour of Rockfill Page 6.68

case studies, particularly those where first filling occurred soon after the end of
construction. For the dumped rockfill case studies the large variation may be
attributable to high rates of leakage and possible collapse settlement of the rockfill
during first filling amongst other factors. The high rate of settlement after the end of
first filling for some case studies is an indication that the response of the rockfill under
changes in stress is not elastic. In addition, higher time dependent settlement rates are
likely under the higher stress conditions imposed at full supply level.
Method (i) is the considered the preferable method to use for prediction of crest
settlements attributable to first filling.

0.6
dumped rockfill
Crest settlement attributable to first

compacted gravels
0.5
very high strength compacted rockfill
filling (% of dam height)

medium to high strength compacted rockfill


0.4

compacted gravels and


0.3 very high strength rockfills

compacted medium to
0.2 high strength rockfills

0.1

0
0 25 50 75 100 125 150 175 200
Embankment Height (m)
Figure 6.39: Crest settlement attributable to first filling (excluding time dependent
effects)

(c) Identification of “abnormal” crest settlement behaviour post construction


Figure 6.31 to Figure 6.36 are useful in providing upper and lower bounds of
deformation behaviour and identification of potential “abnormal” deformation
behaviour. In the case of Scotts Peak dam (Figure 6.35) the post first filling
deformation behaviour is distinctly different from the typical behaviour. In this case
significant leakage occurred during the latter stages of first filling due to cracking of the
asphaltic concrete membrane face (Figure 6.28) resulting in wetting and saturation of
the rockfill. Collapse type settlements occurred on wetting of the medium strength
argillite rockfill, which was prone to strength loss on saturation.
Chapter 6: Deformation Behaviour of Rockfill Page 6.69

6.4.4.2 Post-Construction Internal Settlement


Post-construction internal settlements (under the embankment crest) were analysed for
thirteen of the compacted CFRD case studies, most of which are Australian CFRD. The
settlements were determined from the records of crest monitoring points and internal
settlement gauges and divided into roughly four zones; the upper 25%, the central upper
25%, the central lower 25% and the bottom 25%. The settlements were analysed from
both the start of monitoring (or first filling) and also from the end of first filling. The
findings are summarised in Table 6.11 and indicate:
• For very high strength compacted rockfills and compacted gravels, a very high
percentage of the total settlement from initial monitoring occurs in the bottom 25%
of the embankment and a negligible amount within the upper 25 to 50%.
• For the medium to high strength rockfills this trend is evident for some case studies,
whilst for others the distribution is more evenly spread.
• After the end of first filling the percentage settlement in the bottom 25% of the
embankment decreases (compared to the percent settlement measured from the start
of first filling), typically in the order of 10 to 20%.

The reason for the observed distributions of small settlements in the upper part and
very high settlements in the lower part of the embankment can be explained by a
combination of several factors:
• On first filling the increase in stress levels (above those previously experienced) will
be greatest at the base of the embankment and lowest at the crest.
• The non-linear stress-strain relationship of rockfill, which has a general tendency of
decreasing tangent modulus with increasing vertical stress (Figure 6.17). Therefore,
combined with stress changes from first filling, greater settlements would be
expected in the lower portion of the embankment.
• The tendency for time dependent strain rates to be dependent on stress level.
Greater time dependent strains would therefore be anticipated in the deeper sections
of the embankment.

In Table 6.11 the medium to high strength rockfills have, in most cases, been
sourced from sedimentary rock types, and the very high strength rockfills from igneous
or metamorphic rock types. The sedimentary rock types typically show a greater loss in
Chapter 6: Deformation Behaviour of Rockfill Page 6.70

strength on wetting and therefore, may account for the greater percent of total settlement
in the upper portion of the embankment for several cases due to rainfall infiltration.
Zoning of the embankment will also have an affect as the results for Mangrove
Creek dam indicate. A high proportion of the post-construction crest settlement under
the embankment centreline for Mangrove Creek is from within the random fill zone
(Zone 3B - weathered to fresh sedimentary rockfill) rather than in the better quality
Zone 3A rockfill (fresh siltstones and sandstones).

Table 6.11: Location of internal post-construction vertical settlement in CFRD


CFRD Name Intact *2 Height Start of No of Percent of Total Crest Settlement
Rock (m) Time Years Top 25 to 50 to Bottom
1
Strength Period * Data 25% 50% 75% 25%
Crotty Gravel 83 Start M 2.2 4 6 6 84
Bastyan VH 75 Start FF 6.75 4 18 78
Cethana VH 110 Start ff 14.5 0 51 49
End FF 13.5 (heave) 64 36
Kotmale VH 90 Start M 1.5 (heave) 25 12 63
Murchison VH 94 Start M 11.5 (heave) 27 37 36
End FF 9.25 14 40 30 15
Reece VH 122 Start M 8 4 40 79
End FF 6.25 (heave) 21 49
White Spur H to VH 43 Start M 5.75 40 8 18 33
End FF 4.5 42 4 29 24
Tullabardine H 25 Start M 5 8 52 40
Winneke H 85 Start M 15.5 11 89
End FF 10.5 20 80
Mangrove H 80 Start M 8.3 5 43 26 26
3
Creek *
Mackintosh M to H 75 Start M 7.5 (heave) 23 31 46
End FF 6 (heave) 23 32 45
Serpentine M to H 38 Start M 8 25 42 16 17
End FF 4 15 51 13 21
Scotts Peak M 43 Start M 17 15 7 29 49
End FF 14 37 9 15 39
Notes: *1 M = monitoring; FF = first filling
*2 VH, H and M refer to very high, high and medium intact strength respectively
*3 For Mangrove Creek the central region (25 to 75%) is representative of the random fill zone
(Zone 3B).
Chapter 6: Deformation Behaviour of Rockfill Page 6.71

6.5 DISCUSSION AND R ECOMMENDED M ETHODS FOR P REDICTION


Data on the deformation behaviour of 36 mostly CFRD (35 CFRD and 1 central core
earth and rockfill embankment) has been collected and analysed. The following
discussion provides guidelines for prediction of internal deformations during
construction, face slab deformations during first filling and post-construction crest
settlements of CFRD based on the findings of the analyses undertaken. The proposed
methods are then compared to other existing methods.

6.5.1 GUIDELINES ON DEFORMATION PREDICTION DURING CONSTRUCTION

Empirical methods based on historical performance are useful for prediction of the
deformation of rockfill in CFRD. For relatively standard CFRD designs to be
constructed with quarried rockfills the empirical methods are considered to provide
reasonable estimates of deformation, at least comparable, and possibly superior, to those
based on the results of triaxial and oedometer laboratory tests given the difficulties and
limitations associated with particle size, particle size distribution and sample
preparation of laboratory samples.
Prediction of rockfill deformation (both numerical and empirical) is based on the
quality of representation of the non-linear stress-strain relationship of the rockfill, which
is strongly dependent on the rockfill properties, in particular its intact strength, method
of placement and particle size distribution after compaction. The embankment height is
a significant factor as this predicates the level of applied stress within the embankment.
Valley shape is a significant factor for embankments constructed in narrow valleys due
to the effects of cross-valley arching and resultant reduction in applied stresses. Typical
stress-strain relationships of rockfill from field measurements during construction
(Figure 6.15) show that the relationship is non-linear and has a general trend of
decreasing secant (and tangent) modulus with increasing applied stress.
The proposed method of prediction of a stress-strain relationship for the rockfill is to
firstly derive a representative secant modulus at the end of construction, Erc, from
Figure 6.21 and then apply correction factors for layer thickness, stress level and valley
shape. The representative secant modulus at the end of construction, Erc, is for Zone 3A
rockfill, placed in layers 0.9 to 1.2 m thick, compacted with 4 to 6 passes of a 10 tonne
smooth drum vibratory roller and water added, and applies to average vertical stresses
of (Table 6.7):
Chapter 6: Deformation Behaviour of Rockfill Page 6.72

• 1400 kPa for the very high strength, well-compacted rockfills;


• 800 kPa for the medium to high strength, well-compacted rockfills; and
• 1500 kPa for the well-compacted gravels.

To estimate the secant modulus at the end of construction:


• For the proposed rockfill source, estimate the UCS within the broad ranges of Table
6.1 and the rockfill particle size distribution, most likely from compaction trials.
For existing dams (or those in the process of construction) use the actual average
particle size distribution.
• Estimate the representative Erc from the equations shown on Figure 6.21 using the
estimated D80 value and the appropriate equation for intact strength (either very high
strength or medium to high strength).
• For medium strength well-compacted rockfills apply a correction factor of 0.7 to the
Erc value determined from the equation for medium to high strength rockfills.
• For Zone 3B or other thicker layer rockfills comprising similar rock type (and UCS
range) to Zone 3A, apply a correction factor of 0.5 to obtain a representative Erc for
rockfill placed in 1.5 to 2.0 m layers. This correction factor is based on the ratio of
Erc (of Zone 3A to Zone 3B) from six field cases.
• To account for the non-linearity of the stress-strain relationship for rockfill,
estimates of Erc at stress levels less than or greater than that representative for the
corrected Erc can be estimated from historical records (Figure 6.16). Although the
relationship is likely to be exponential, the data (Figure 6.16) indicates a linear
approximation is reasonable over a defined vertical stress range. The following
corrections are suggested:
− For very high strength rockfills apply a linear correction of ±7.5% per 200 kPa
to the E rc estimated from Figure 6.21 from a vertical stress of 1400 kPa
(applicable range is 400 to 1600 kPa). Apply positive corrections for decreasing
stresses and negative corrections for increasing stresses.
− For medium to high strength rockfills apply a linear correction of ±6 % per 200
kPa to the Erc estimated from Figure 6.21 from a vertical stress of 800 kPa
(applicable range is 200 to 1200 kPa).
• For one-dimensional analyses appropriate correction factors for stress reduction due
to the shape of the embankment should be considered. These can be estimated from
equations and charts such as those published by Poulos et al (1972).
Chapter 6: Deformation Behaviour of Rockfill Page 6.73

• To correct for valley shape either:


− Model the valley and the dam using 3-dimensional numerical analysis, and a
stress-strain relationship for the rockfill based on the Erc versus vertical stress
relationship derived above; or
− For simplified one-dimensional analyses, to estimate the maximum settlement
under the centreline in the centre of the valley correct the vertical stress profile
for embankment shape (refer Section 6.2.3) and then correct for valley shape
using the appropriate stress reduction factors given in Table 6.12.

Table 6.12: Approximate stress reduction factors to account for valley shape
Stress Reduction Factor (embankment location)
Wr/H Ratio Average
(river width Abutment Slope Base Mid to Low Mid Upper
(0 to 20%) (20 to 40%) (40 to 65%) (65% to
to height) Angle (degrees)
crest)
0.2 10 to 20 0.93 0.95 0.97 1.0
20 to 30 0.88 0.92 0.96 0.98
30 to 40 0.82 0.88 0.94 0.97
40 to 50 0.74 0.83 0.91 0.96
50 to 60 0.66 0.76 0.86 0.94
60 to 70 0.57 0.69 0.82 0.92
0.5 < 25 1.0 1.0 1.0 1.0
25 to 40 0.93 0.95 0.97 1.0
40 to 50 0.91 0.92 0.95 0.95 – 1.0
50 to 60 0.87 0.88 0.93 0.95 – 1.0
60 to 70 0.83 0.85 0.90 0.95 – 1.0
1.0 All slopes 0.95 - 1.0 0.95 - 1.0 1.0 1.0
Note: Wr = river width H = embankment height

Other relevant issues relating to the estimation of secant modulus and derivation of the
correlation between secant modulus and D80 (Figure 6.21) are:
• Cu, the uniformity coefficient for the particle size distribution curve, is implicitly
allowed for in the D80 value. Generally, decreasing Cu is observed for increasing
D80.
• For materials placed with larger rollers (i.e. 13 to 15 tonne deadweight vibrating
rollers) the data does not indicate an increase in moduli for the greater compactive
effort. This may be due to greater material breakdown under the heavier rollers, and
a resultant reduction of D80, which can then be accounted for directly in Figure 6.21.
Chapter 6: Deformation Behaviour of Rockfill Page 6.74

• For weathered rockfills intact strength will be lower than for fresh rock, which will
result in a decrease in secant moduli, but the greater breakdown will give an
increase in moduli associated with a reduction in D80.
• The E rc values in Figure 6.21 for each data point have been determined from
published internal deformation records, and generally from the lower 60% of the
embankment and within the central portion of this lower section. The Erc values
have also been corrected for cross-valley arching effects using the normalised
vertical stress charts (Figure 6.19). It is recognised that these may not be a true
representation of the actual stresses due to assumptions made in the modelling
process (refer Section 6.4.2). However, errors associated with the assumptions in
modelling are minimised by applying the same assumptions for prediction as is
proposed in the above.
• If testing on the proposed rock source for rockfill indicates a significant reduction in
UCS on wetting is likely and only limited water has been, or is proposed to be, used
in construction, then this strength reduction should be taken into consideration.
There is not sufficient data to advise on how this should be done; however, the
rockfill is likely to be more susceptible to collapse settlements on wetting due to
rainfall and flooding.

Limitations of the data and application of the proposed methodology include:


• For well-compacted gravels only a limited number of cases were available,
insufficient to undertake analysis. These materials generally have significantly
greater secant moduli at any given stress level in comparison to quarried, very high
strength rockfill. Where the rockfill is of finer size (e.g. Crotty Zone 3A) the use of
large scale laboratory test that virtually encompass the field particle size distribution
would provide suitable estimates of moduli provided the layering and density in the
field are reflected in the laboratory testing.
• For the data presented in Figure 6.15 (vertical stress versus vertical strain), the
stress-strain curves do not pass through the origin. This is because zero strain is
assumed at a vertical stress equivalent to the mid-height between hydrostatic
settlement gauges (HSGs). Zero strain is generally at vertical stresses in the range
50 to 200 kPa (depending on the vertical distance between HSGs).
• During the actual embankment construction, monitoring should be undertaken as a
check on the pre-construction estimation of deformation. Deformation estimates
Chapter 6: Deformation Behaviour of Rockfill Page 6.75

can be reassessed, if required, as construction proceeds using the stress-strain


relationship/s extrapolated from the actual field records.

For estimating deformations due to embankment raising, it is recommended that the


modulus be obtained from monitoring during the earlier construction phase if that is
available. If not, the methods outlined above can be used to estimate the secant
modulus versus vertical stress relationship. Simplistic estimates of deformation could
be obtained by modelling the embankment in a series of layers of varying tangent
moduli (estimated from the secant moduli vertical stress relationship). For a more
rigorous analysis numerical modelling would be appropriate.
For dumped rockfill there is not sufficient data from which to give guidelines on
moduli during construction. The data records from the limited number of cases is
summarised in Section 6.4.2.6.

6.5.2 GUIDELINES ON DEFORMATION PREDICTION DURING FIRST FILLING

As discussed in Section 6.4.3, the modulus on first filling, Erf, estimated using the
method by Fitzpatrick et al (1985) is an artefact of the method used to calculate it and is
not a true modulus of the rockfill, as Fitzpatrick et al (1985) recognised. The high
values of Erf compared to Erc are to be expected given the stress paths on first filling in
the upstream slope (Figure 6.23b and Figure 6.24b). The stress paths show an
unloading then reloading of deviatoric stress whilst compressive stresses continue to
increase with increasing reservoir level.
The predictive method of face slab deformation comprises estimation of the ratio
Erf / Erc from Figure 6.22. Given that E rc is already known or estimated, E rf can be
estimated. The face slab deflection can then be predicted using Equation 6.2 after
Fitzgerald et al (1985) (refer Figure 6.1b). The method is considered applicable to the
region of the face slab between 20% and 60% of the embankment height, the region of
maximum face slab deflection on first filling in most cases.
Guidelines for use of Figure 6.22 for estimation of Erf are:
• Estimate the ratio Erf / Erc based on the embankment height and upstream slope
angle. Trendlines are given for upstream slopes of 1.3 – 1.4H to 1V, and 1.5H to
1V, which cover most CFRD designs.
Chapter 6: Deformation Behaviour of Rockfill Page 6.76

• Estimate Erc, the representative secant moduli at end of construction, of the Zone 3A
rockfill from the methods outlined in Section 6.5.1. The E rc value should be
adjusted for vertical stress such that it is representative of the lower 50% of the
rockfill in the central region of the embankment. The Erc values used in the
derivation of Figure 6.22 were not corrected for arching effects due to valley shape
because valley shape is potentially likely to affect both Erc and Erf, and would
therefore be taken into consideration in the Erf / Erc ratio. Hence, Erc estimates
derived from Figure 6.21 (corrected for valley shape effects) must be un-corrected.
• Erf can then be estimated by multiplication of the Erf / Erc ratio by Erc (not corrected
for valley shape).

In using Figure 6.22 for estimating the Erf / Erc ratio, consideration should be given
to the qualifications to this figure as discussed in Section 6.4.3. The trendline for
upstream slope angles of 1.5H to 1V is based on a limited number of cases and it should
therefore be recognised that the estimated Erf / Erc ratio will be very approximate.
The zoning geometry of the embankment has a significant effect on the deformation
behaviour of the face slab on first filling (Section 6.4.3) and the method of prediction of
the face slab deformation is dependent on this geometry and consideration of the
differential moduli between the main rockfill zones. The empirical predictive method
developed (based on Figure 6.22) is appropriate for CFRD with relatively simple zoning
geometries comprising a significant Zone 3A (at least 50 to 60% of the main rockfill)
and/or where large variations in rockfill moduli are not likely between rockfill zones.
For embankments with complex geometries (particularly in the upstream half of the
embankment) and/or large moduli variations between the rockfill zones, finite element
analyses should be considered. The detailed analysis would be justified considering the
potential for face slab cracking as a result of possible development of tensile stresses.

6.5.3 GUIDELINES ON DEFORMATION PREDICTION POST-CONSTRUCTION

Guidelines for the prediction of post-construction crest settlements are given in Section
6.4.4.1.4. The methods proposed are considered to provide an improvement on current
available methods of crest settlement prediction for CFRD.
Chapter 6: Deformation Behaviour of Rockfill Page 6.77

The method of prediction comprises separate estimation of the time dependent


deformation component and the crest settlement attributable to first filling, and then
summation to estimate the total post construction settlement. In summary:
• The time dependent crest settlement rate (as a percentage of the embankment height
per log cycle of time) for compacted rockfill can be estimated from Figure 6.38.
The estimate is based on the embankment height and intact strength of the
compacted rockfill under the embankment centreline. The zero time, to, has been
taken at the end of the main rockfill construction.
• For dumped rockfills, estimates of the long-term crest settlement rate (as a
percentage of the embankment height per log cycle of time) can be derived from
either Figure 6.31 or Table 6.10 based on the time period after the end of main
rockfill construction.
• Two methods are provided for prediction of the crest settlement during or
attributable to first filling. Method (i) or Figure 6.39, the preferred method, is an
estimation of the crest settlement attributable to first filling based on the
embankment height, intact rock strength and placement method. The degree of
variability is relatively large, particularly for the dumped rockfill cases, due to the
number of factors influencing the deformation behaviour.
• The time dependent deformations and deformations on first filling occur
concurrently.

The plots of the historical records of case studies (Figure 6.31 to Figure 6.36) are
useful in providing upper and lower bounds of deformation behaviour and identification
of potential “abnormal” deformation behaviour based on the method of placement and
intact strength of compacted rockfill. As discussed in Section 6.4.4.1 the plots for
settlements after the end of main rockfill construction (Figure 6.31 to Figure 6.33)
should be used with caution due to the influences of first filling and initial time of
reading.

6.6 CONCLUSIONS
The proposed empirical predictive methods for estimation of the deformation of rockfill
(summarised in Section 6.5), specifically targeted at the deformation of CFRD, are
considered to provide improvements over currently available empirical predictive
Chapter 6: Deformation Behaviour of Rockfill Page 6.78

methods because of the larger good quality data set available, and consideration of
significant factors which influence the deformation behaviour. The methods are based
on historical records of embankment performance and allow for estimation/prediction of
the vertical deformation during construction, deformation of the face slab and post-
construction crest settlement. Where possible the significant factors affecting the
deformation behaviour of rockfill, evaluated from the analysis of historical records,
were incorporated into the empirical predictive methods.
The results of the historical data presented on the stress-strain behaviour of rockfill
measured in the field is considered to have advantages over rockfill properties estimated
by laboratory test methods. Laboratory testing of rockfill has the disadvantages of
limitation on particle size and placement methods, both of which are shown to have a
significant effect on the prediction of the rockfill modulus. Methods, although
relatively crude, are proposed for estimation of the stress-strain relationship of a
compacted rockfill sample and could therefore be used as a basis for the constitutive
model in finite element analyses.
The proposed empirical predictive methods of deformation are generally applicable
to the more typical CFRD designs. For relatively complex geometries, particularly
those that incorporate high differential moduli between rockfill zones (e.g. compacted
gravels and rockfills such as Aguamilpa dam) numerical methods are considered more
appropriate. They are considered particularly appropriate for modelling potential tensile
stress development in the concrete face that could result in cracking.
Chapter 7: Deformation behaviour of embankment dams Page i

TABLE OF CONTENTS

7.0 THE DEFORMATION BEHAVIOUR OF


EMBANKMENT DAMS............................................................. 7.1
7.1 INTRODUCTION TO THIS C HAPTER ................................................................... 7.1
7.2 LITERATURE R EVIEW ...................................................................................... 7.2
7.2.1 Failure and Accident Statistics.................................................................. 7.2
7.2.2 Historical Development of Embankment Dams as this Affects
Deformation Behaviour.............................................................................7.3
7.2.3 Factors Affecting The Deformation of Embankment Dams ...................... 7.5
7.2.4 Predictive Methods of Deformation Behaviour.........................................7.9
7.3 DATABASE OF CASE S TUDIES ........................................................................7.17
7.3.1 Earthfill and Zoned Earth and Earth-Rockfill Embankments.................7.17
7.3.2 Puddle Core Earthfill Embankments....................................................... 7.18
7.4 GENERAL DEFORMATION BEHAVIOUR DURING CONSTRUCTION OF
EARTHFILL AND ZONED E ARTH AND EARTH-ROCKFILL EMBANKMENTS ...... 7.18
7.4.1 Stresses During Construction.................................................................. 7.23
7.4.2 Lateral Deformation of the Core During Construction .......................... 7.43
7.4.3 Vertical Deformation of the Core During Construction ......................... 7.51
7.5 GENERAL DEFORMATION BEHAVIOUR ON FIRST FILLING OF EARTHFILL
AND ZONED EARTH AND E ARTH-ROCKFILL EMBANKMENTS ......................... 7.69
7.5.1 Effect of Water Load on the Core............................................................ 7.70
7.5.2 Collapse Compression During First Filling ........................................... 7.73
7.5.3 Lateral Surface Deformation Normal to the Dam Axis During First
Filling. .....................................................................................................7.84
7.6 GENERAL POST CONSTRUCTION DEFORMATION BEHAVIOUR OF
EARTHFILL AND ZONED E ARTH AND EARTH-ROCKFILL EMBANKMENTS ...... 7.99
7.6.1 Post Construction Internal Vertical Deformation of the Core..............7.100
7.6.2 Post Construction Total Deformation of Surface Monitoring Points ...7.102
7.6.3 Post Construction Crest Settlement Versus Time..................................7.110
7.6.4 Post Construction Horizontal Displacement of the Crest Normal to
the Dam Axis. ........................................................................................7.132
7.6.5 Post Construction Deformation of the Mid to Upper Downstream
Slope ......................................................................................................7.140
Chapter 7: Deformation behaviour of embankment dams Page ii

7.6.6 Post Construction Deformation of the Upper Upstream Slope and


Upstream Crest......................................................................................7.148
7.7 GENERAL DEFORMATION BEHAVIOUR OF PUDDLE CORE EARTHFILL
EMBANKMENTS ...........................................................................................7.158
7.7.1 Deformation During Construction of Puddle Core Earthfill
Embankments.........................................................................................7.158
7.7.2 Deformation During First Filling of Puddle Core Earthfill
Embankments.........................................................................................7.159
7.7.3 Deformation Behaviour Post First Filling of Puddle Core Earthfill
Dams......................................................................................................7.166
7.8 “ABNORMAL” E MBANKMENT DEFORMATION BEHAVIOUR – METHODS OF
IDENTIFICATION ..........................................................................................7.185
7.9 “ABNORMAL” DEFORMATION BEHAVIOUR DURING CONSTRUCTION OF
EARTH AND EARTH-ROCKFILL EMBANKMENTS ..........................................7.187
7.9.1 Plastic Deformation of the Core During Construction.........................7.188
7.9.2 Collapse Compression of the Central Earthfill Zone............................7.191
7.9.3 Reservoir Filling During Construction .................................................7.192
7.9.4 Shear Surface Development in the Core During Construction.............7.193
7.10 “ABNORMAL” DEFORMATION BEHAVIOUR POST CONSTRUCTION OF
ZONED E ARTH AND ROCKFILL EMBANKMENTS ...........................................7.196
7.10.1 Rockfill Susceptibility to Collapse Compression...................................7.197
7.10.2 Deformation on First Filling in Embankments where Collapse
Compression occurs in the Upstream Rockfill Shoulder.......................7.201
7.10.3 Development of Shear Surfaces Within the Earthfill Core....................7.217
7.10.4 Post First Filling Acceleration in Deformation that is Not Known to
be Shear Related....................................................................................7.241
7.10.5 Other Case Studies with Potentially “Abnormal” Deformation
Behaviour Post Construction ................................................................7.243
7.11 “ABNORMAL” DEFORMATION BEHAVIOUR POST CONSTRUCTION OF
EARTHFILL AND ZONED EMBANKMENTS WITH VERY B ROAD CORE
WIDTHS .......................................................................................................7.252
7.11.1 Collapse Compression of the Earthfill on Wetting................................7.252
7.11.2 “High” Shear Stress or Marginal Stability Conditions Within the
Embankment ..........................................................................................7.260
7.11.3 The Effect of the Development of the Phreatic Surface on the
Displacement of the Crest and Downstream Shoulder. ........................7.269
7.11.4 Other Cases of Potential “Abnormal” Deformation Behaviour...........7.270
Chapter 7: Deformation behaviour of embankment dams Page iii

7.12 “ABNORMAL” DEFORMATION BEHAVIOUR OF PUDDLE CORE E ARTHFILL


DAMS ..........................................................................................................7.274
7.12.1 Comparison with the Deformation Behaviour of Other Puddle Dams .7.275
7.12.2 General Movement Trends Indicative of Deformation to Failure.........7.278
7.12.3 Other Indicators of “Abnormal” Deformation Behaviour....................7.279
7.13 SUMMARY OF “ABNORMAL” DEFORMATION BEHAVIOUR ...........................7.280
7.14 SUMMARY AND METHODS FOR PREDICTION OF DEFORMATION OF
EMBANKMENT DAMS ..................................................................................7.287
7.14.1 Earthfill, and Zoned Earth and Earth-Rockfill Dams ...........................7.288
7.14.2 Prediction of Deformation Behaviour During Construction for
Earthfill and Zoned Earth and Earth-Rockfill Embankments...............7.288
7.14.3 Prediction of Deformation Behaviour Post Construction for Earthfill
and Zoned Earth and Earth-Rockfill Embankments..............................7.291
7.14.4 Puddle Core Earthfill Dams..................................................................7.300
7.15 CONCLUSIONS .............................................................................................7.302
Chapter 7: Deformation behaviour of embankment dams Page iv

LIST OF TABLES
Table 7.1: Mechanisms affecting the long-term deformation behaviour of old
puddle core earthfill embankments (Tedd et al 1997a)........................................... 7.8
Table 7.2: Vertical compression of rolled, well-compacted earthfills measured
during construction (adapted from Sherard et al 1963)......................................... 7.12
Table 7.3: Published ranges of post construction deformation of embankment
dams ...................................................................................................................... 7.14
Table 7.4: Central core earth and rockfill embankments in the database .................... 7.19
Table 7.5: Zoned earth and rockfill embankment case studies .................................... 7.20
Table 7.6: Zoned earthfill embankment case studies ................................................... 7.21
Table 7.7: Earthfill embankment case studies.............................................................. 7.21
Table 7.8: Puddle core earthfill embankment case studies .......................................... 7.22
Table 7.9: Multiplying coefficients for the calculation of the total vertical stress at
the embankment centreline at end of construction (Equation 7.3)........................ 7.26
Table 7.10: Properties of central core used in finite difference analyses..................... 7.29
Table 7.11: Lateral deformations of the central core at end of construction for
central core earth and rockfill embankments. ....................................................... 7.44
Table 7.12: Summary of embankment and earthfill properties for cases used in the
analysis of lateral core deformation during construction. ..................................... 7.46
Table 7.13: Confined secant moduli during construction for well-compacted, dry
placed earthfills. .................................................................................................... 7.57
Table 7.14: Equations of best fit for core settlement versus embankment height
during construction................................................................................................ 7.68
Table 7.15: Equations of best fit for core settlement (as a percentage of
embankment height) versus embankment height during construction.................. 7.68
Table 7.16: Properties of central core used in finite difference analyses..................... 7.80
Table 7.17: Lateral displacement of the crest (centre to downstream edge) on first
filling. .................................................................................................................... 7.89
Table 7.18: Thirteen cases with largest downstream crest displacement on first
filling ..................................................................................................................... 7.92
Table 7.19: Typical range of post construction crest settlement. ...............................7.104
Table 7.20: Typical range of post construction settlement of the upstream and
downstream shoulders.........................................................................................7.109
Table 7.21: Summary of long-term crest settlement rates (% per log cycle of
time). ...................................................................................................................7.122
Table 7.22: Range of long-term settlement rate of the downstream shoulder. ..........7.141
Chapter 7: Deformation behaviour of embankment dams Page v

Table 7.23: Range of long-term settlement rate of the upper upstream slope and
upstream crest region. .........................................................................................7.151
Table 7.24: Summary of the post construction surface deformations of the puddle
core earthfill dam case studies. ...........................................................................7.168
Table 7.25: Embankments for which collapse compression caused moderate to
large settlements..................................................................................................7.199
Table 7.26: Figure references for post construction crest deformation......................7.292
Table 7.27: Figure references for post construction deformation of the
embankment shoulders........................................................................................7.293
Table 7.28: References to post construction settlement magnitude tables and plots .7.295
Table 7.29: Embankment crest region, typical range of post construction
settlement and long-term settlement rate ............................................................7.295
Table 7.30: Embankment shoulder regions, typical range of post construction
settlement and long-term settlement rate ............................................................7.296
Table 7.31: References to tables and figures of long-term settlement rate ................7.296
Table 7.32: Predictive methods of long-term crest settlement and displacement
under normal reservoir operating conditions for puddle core earthfill
embankments.......................................................................................................7.301
Chapter 7: Deformation behaviour of embankment dams Page vi

LIST OF FIGURES
Figure 7.1: Embankment construction indicating (a) broad layer width to
embankment depth ratio in the early stages of construction and (b) narrow
layer width to embankment depth ratio in the latter stages of construction.......... 7.24
Figure 7.2: (a) Total vertical stress (σz,EOC ) contours at end of construction for
1.8H to 1V embankment slopes, and (b) vertical stress ratio, Z1, versus h/H
under the embankment axis for various embankment slopes................................ 7.26
Figure 7.3: Embankment model for finite difference analysis during construction.....7.29
Figure 7.4: Total vertical stress, total lateral stress, and lateral (horizontal)
displacement distributions at end of construction for zoned earthfill finite
difference analysis................................................................................................. 7.30
Figure 7.5: Finite difference modelling results; (a) total vertical stress on the dam
centreline at end of construction, (b) settlement profile at dam axis at end of
construction, and (c) lateral deformation at Point a (Figure 7.3) during
construction. .......................................................................................................... 7.32
Figure 7.6: Finite element analysis of zoned embankments highlighting arching of
core zone from (a) Beliche dam at about mid-height, elevation 22.5 m
(Naylor et al 1997), and (b) a puddle core embankment (Dounias et al 1996).....7.33
Figure 7.7: Estimated versus measured (from pressure cells) total vertical stresses
during construction................................................................................................ 7.34
Figure 7.8: Total vertical stress profiles from the numerical analysis of
construction for zoned embankments with different core slopes; (a) 0.2H to
1V case, (b) 0.5H to 1V case, and (c) total vertical stress at h/H = 0.70 with
increasing embankment height. ............................................................................. 7.36
Figure 7.9: Idealised types of pore water pressure response in the core during
construction. .......................................................................................................... 7.38
Figure 7.10: Pore water pressure response during construction in the core zone of
clayey earthfills placed at close to or wet of Standard optimum moisture
content. .................................................................................................................. 7.39
Figure 7.11: Triaxial isotropic compression tests with staged undrained loading
and partial drainage on a partially saturated silty sand (Bishop 1957). ................ 7.41
Figure 7.12: Estimated lateral displacement ratio of the core at end of
construction. .......................................................................................................... 7.45
Figure 7.13: Estimated lateral displacement ratio of the core versus fill height
above gauge during construction........................................................................... 7.45
Chapter 7: Deformation behaviour of embankment dams Page vii

Figure 7.14: Lateral displacement ratio and pore water pressure versus fill height
or measured total vertical stress at (a) Dartmouth dam; (b) Blowering dam;
(c) Talbingo dam and (d) Thomson dam............................................................... 7.50
Figure 7.15: Vertical stress versus strain of the core during construction for
selected cross-arm intervals of selected case studies; (a) dry placed clay
cores, and (b) dry placed dominantly sandy and gravelly cores with plastic
fines. ...................................................................................................................... 7.54
Figure 7.16: Confined secant moduli of the core versus vertical stress during
construction for selected cross-arm intervals of selected case studies, (a) dry
placed clay cores, and (b) dry placed dominantly sandy and gravelly cores
with plastic fines. .................................................................................................. 7.55
Figure 7.17: Seating settlements at low stresses suspected as cause of low moduli
estimates at low stresses........................................................................................ 7.56
Figure 7.18: Vertical strain versus effective vertical stress in the core at end of
construction for (a) dry placed clay earthfills and (b) dry placed dominantly
sandy and gravelly earthfills. ................................................................................ 7.58
Figure 7.19: Confined secant moduli versus effective vertical stress in the core at
end of construction for (a) dry placed clay earthfills and (b) dry placed
dominantly sandy and gravelly earthfills.............................................................. 7.59
Figure 7.20: Vertical strain in the core versus fill height during construction at
selected cross-arm intervals in selected case studies of earthfills placed close
to or wet of Standard optimum, for (a) thin core (combined slopes less than
0.5H to 1V), and (b) medium cores (combined slope ≥ 0.5H to 1V and < 1H
to 1V). ................................................................................................................... 7.62
Figure 7.21: Vertical strain in the core versus fill height at end of construction for
dominantly sandy and gravelly earthfills with plastic fines placed close to or
wet of Standard optimum, for (a) thin cores (combined slopes less than 0.5H
to 1V), and (b) medium to thick cores (combined slope ≥ 0.5H to 1V). .............. 7.63
Figure 7.22: Vertical strain of the core versus fill height at end of construction for
clay earthfills placed close to or wet of Standard optimum moisture content,
for (a) thin core (combined slopes less than 0.5H to 1V), and (b) medium to
thick cores (combined slope ≥ 0.5H to 1V). ......................................................... 7.64
Figure 7.23: Core settlement during construction of earth and earth-rockfill
embankments (a) including Nurek dam, and (b) excluding Nurek dam. ..............7.67
Figure 7.24: Effect of reservoir filling on a zoned embankment (Nobari and
Duncan 1972b) ...................................................................................................... 7.70
Figure 7.25: Water load acting on a homogeneous earthfill dam. ...............................7.71
Chapter 7: Deformation behaviour of embankment dams Page viii

Figure 7.26: Lateral stress distribution at (a) end of construction, and (b) reservoir
full condition, for central core earth and rockfill embankment with core of
similar compressibility properties to the rockfill. ................................................. 7.72
Figure 7.27: Lateral stress distribution at (a) end of construction, and (b) reservoir
full condition, for a central core earth and rockfill embankment with wet
placed core of low undrained strength. ................................................................. 7.73
Figure 7.28: Compression curves for dry and wet states and collapse compression
from the dry to wet state for Pyramid gravel in the laboratory oedometer test
(Nobari and Duncan 1972a). ................................................................................. 7.75
Figure 7.29: Idealised effect of matric suction on collapse compression of
earthfills (Khalili 2002)......................................................................................... 7.75
Figure 7.30: Cracking caused by post construction differential settlement between
the core and the dumped rockfill shoulders (Sherard et al 1963).......................... 7.77
Figure 7.31: Embankment model for finite difference analysis during first filling. .... 7.78
Figure 7.32: Model of the stress-strain relationship of the upstream rockfill
incorporating collapse compression (in one-dimensional vertical
compression). ........................................................................................................ 7.80
Figure 7.33: Case 8 – “dry” placed, very stiff core modelling collapse
compression in the upstream rockfill; numerical results of (a) vertical and (b)
lateral displacement on first filling........................................................................ 7.82
Figure 7.34: Case 9 – “dry” placed, very stiff core without collapse compression;
numerical results of (a) vertical and (b) lateral displacement on first filling........ 7.82
Figure 7.35: Case 10 – “wet” placed clay core of low undrained strength
modelling collapse compression in the upstream rockfill. Numerical results
of vertical and lateral deformation on first filling; (a) and (b) for reservoir at
40% of embankment height, (c) and (d) for reservoir at 98% of embankment
height. .................................................................................................................... 7.83
Figure 7.36: Case 11 – “wet” placed clay core without collapse compression;
numerical results of (a) vertical and (b) lateral displacement on first filling........7.84
Figure 7.37: Regional division of the embankment for analysis of post
construction surface deformation. ......................................................................... 7.85
Figure 7.38: Lateral displacement of the crest (centre or downstream region) on
first filling versus embankment height (displacement is after the end of
embankment construction). ................................................................................... 7.88
Figure 7.39: Lateral displacement of the downstream slope (mid to upper region)
on first filling versus embankment height (displacement is after the end of
embankment construction). ................................................................................... 7.95
Chapter 7: Deformation behaviour of embankment dams Page ix

Figure 7.40: Lateral displacement at the upper upstream slope of upstream crest
region on first filling versus embankment height (displacement after the end
of embankment construction)................................................................................ 7.97
Figure 7.41: Lateral displacement of the crest (centre to downstream region) on
first filling versus downstream slope. ................................................................... 7.98
Figure 7.42: Lateral displacement of the downstream slope (mid to upper region)
on first filling versus downstream slope. .............................................................. 7.98
Figure 7.43: Post construction internal settlement of the core at Talbingo dam
(IVM ES1 at the main section)............................................................................7.101
Figure 7.44: Post construction internal settlement of the core at Copeton dam
(IVM B, 9 m downstream of dam axis at main section). ....................................7.101
Figure 7.45: Post construction internal settlement of the core at Bellfield dam. .......7.102
Figure 7.46: Post construction crest settlement at 3 years after end of construction,
(a) all data, (b) data excluding Ataturk. ..............................................................7.105
Figure 7.47: Post construction crest settlement at 10 years after end of
construction. ........................................................................................................7.105
Figure 7.48: Post construction crest settlement at 20 to 25 years after end of
construction. ........................................................................................................7.106
Figure 7.49: Post construction settlement of the downstream slope (mid to upper
region) at 3 years after end of construction. ........................................................7.107
Figure 7.50: Post construction settlement of the downstream slope (mid to upper
region) at 10 years after end of construction. ......................................................7.107
Figure 7.51: Post construction settlement of the upper upstream slope and
upstream crest region at 3 years after end of construction. .................................7.108
Figure 7.52: Post construction settlement of the upper upstream slope and
upstream crest region at 10 years after end of construction. ...............................7.108
Figure 7.53: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey earthfills; (a) all data,
(b) data excluding Ataturk. .................................................................................7.113
Figure 7.54: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey earthfills.......................7.114
Figure 7.55: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey sand to clayey gravel
(SC to GC) earthfills. ..........................................................................................7.114
Figure 7.56: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey sand to clayey gravel
(SC to GC) earthfills. ..........................................................................................7.115
Chapter 7: Deformation behaviour of embankment dams Page x

Figure 7.57: Crest settlement versus time for zoned embankments with thin to
medium width central core zones of silty sand to silty gravel (SM to GM)
earthfills...............................................................................................................7.116
Figure 7.58: Crest settlement versus time for zoned embankments with central
core zones of clayey earthfills of thick width (1 to 2.5H to 1V combined
width). .................................................................................................................7.116
Figure 7.59: Crest settlement versus time for zoned embankments with central
core zones of silty to clayey gravel and sand (SC, GC, SM, GM) earthfills of
thick width (1 to 2.5H to 1V combined width). ..................................................7.117
Figure 7.60: Crest settlement versus time for embankments with very broad core
widths (> 2.5H to 1V combined width) and limited foundation influence. ........7.118
Figure 7.61: Crest settlement versus time for embankments with very broad core
widths (> 2.5H to 1V combined width) and potentially significant foundation
influence..............................................................................................................7.118
Figure 7.62: Long-term post construction crest settlement rate for zoned earthfill
and earth-rockfill embankments of thin to thick core widths, (a) all data, and
(b) data excluding Ataturk. .................................................................................7.120
Figure 7.63: Long-term post construction crest settlement rate for embankments of
very broad core width, (a) all data, and (b) data excluding Belle Fourche,
Roxo and Rector Creek. ......................................................................................7.121
Figure 7.64: Crest settlement versus time for central core earth and rockfill
embankments with central core zones of clayey earthfills and steady post first
filling reservoir operation....................................................................................7.127
Figure 7.65: Crest settlement versus time for central core earth and rockfill
embankments with thin to thick central core zones of clayey earthfills
subjected to seasonal reservoir fluctuation..........................................................7.127
Figure 7.66: Post construction settlement and reservoir operation at Enders dam
(data courtesy of USBR). ....................................................................................7.131
Figure 7.67: Post construction settlement at main section and reservoir operation
at San Luis dam (data courtesy of USBR). .........................................................7.132
Figure 7.68: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey earthfills.......................7.133
Figure 7.69: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey earthfills.......................7.134
Figure 7.70: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey sand and clayey gravel
(SC/GC) earthfills. ..............................................................................................7.134
Chapter 7: Deformation behaviour of embankment dams Page xi

Figure 7.71: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey sand and clayey gravel
(SC/GC) earthfills. ..............................................................................................7.135
Figure 7.72: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of silty sand and silty gravel (SM/GM)
earthfills...............................................................................................................7.135
Figure 7.73: Crest displacement versus time for zoned embankments with central
core zones of clayey earthfills of thick width (1 to 2.5H to 1V combined
width). .................................................................................................................7.136
Figure 7.74: Crest displacement versus time for zoned embankments with central
core zones of silty to clayey sand and gravel earthfills of thick width. ..............7.136
Figure 7.75: Crest displacement versus time for embankments with very broad
core zones (> 2.5H to 1V width) and with limited foundation influence............7.137
Figure 7.76: Crest displacement versus time for embankments with very broad
core zones (> 2.5H to 1V width) and with potentially significant foundation
influence on the embankment deformation. ........................................................7.137
Figure 7.77: Crest displacement normal to dam axis post first filling.......................7.139
Figure 7.78: Long-term settlement rates for the downstream slope (mid to upper
region) versus embankment height; (a) all data, and (b) data excluding Belle
Fourche................................................................................................................7.143
Figure 7.79: Post construction settlement of the downstream shoulder (mid to
upper region) for selected case studies................................................................7.144
Figure 7.80: Post construction displacement of the downstream shoulder (mid to
upper region) for selected case studies of embankments with very broad core
widths. .................................................................................................................7.145
Figure 7.81: Post construction displacement of the downstream shoulder (mid to
upper region) for selected case studies of central core earth and rockfill
embankments.......................................................................................................7.147
Figure 7.82: Post construction displacement of the downstream shoulder (mid to
upper region) for embankments with compacted earthfills and gravels in the
downstream shoulder. .........................................................................................7.148
Figure 7.83: Long-term settlement rates of the upper upstream slope to upstream
crest region of earthfill and earth-rockfill embankments. ...................................7.152
Figure 7.84: Post construction settlement of the upper upstream slope to upstream
crest region for embankments with poorly compacted rockfill in the upstream
slope. ...................................................................................................................7.153
Figure 7.85: Post construction settlement of the upper upstream slope to upstream
crest region for selected case studies (excluding poorly compacted rockfills)...7.153
Chapter 7: Deformation behaviour of embankment dams Page xii

Figure 7.86: Post construction lateral displacement of the upper upstream slope
and upstream crest region for selected case studies of embankments with
rockfill in the upstream shoulder.........................................................................7.156
Figure 7.87: Post construction lateral displacement of the upper upstream slope to
upstream crest region for selected case studies of embankment with earthfills
and gravels in the upstream shoulder..................................................................7.157
Figure 7.88: Selset Dam internal settlements during construction (data from
Bishop and Vaughan 1962).................................................................................7.159
Figure 7.89: “Idealised” model of collapse compression of poorly compacted
shoulder fill of puddle core earthfill dam on first filling.....................................7.164
Figure 7.90: “Idealised” model of yielding of poorly compacted shoulder fill and
puddle core on first drawdown of a puddle core earthfill dam. ..........................7.164
Figure 7.91: “Idealised” model of collapse compression on wetting of poorly
compacted earthfill in the downstream shoulder of a puddle core earthfill
dam. .....................................................................................................................7.165
Figure 7.92: Post construction crest settlement versus log time of puddle core
earthfill dams.......................................................................................................7.167
Figure 7.93: Definitions of “steady state” conditions during normal reservoir
operating conditions. ...........................................................................................7.172
Figure 7.94: Fluctuation of pore water pressures adjacent to and upstream of the
puddle core with fluctuations in reservoir level..................................................7.174
Figure 7.95: Ramsden Dam, internal horizontal deformation of the puddle core,
1988 to 1990 (Holton 1992)................................................................................7.176
Figure 7.96: Ramsden Dam, internal vertical strains measured in the puddle core
from 1988 to 1990 (data from Tedd et al 1997b and Kovacevic et al 1997). .....7.177
Figure 7.97: Vertical strain (or settlement) at the crest versus drawdown depth for
specific drawdown and refilling cycles (Tedd et al 1997b). ...............................7.179
Figure 7.98: Permanent vertical crest strains during cyclic drawdown (in normal
operating range) for puddle core dams with permeable upstream filling (after
Tedd et al 1997b). ...............................................................................................7.179
Figure 7.99: Ramsden Dam, settlement of SMPs versus time from 1988 to 1990
(data from Tedd et al 1990).................................................................................7.180
Figure 7.100: Ogden Dam, crest settlement versus time (Tedd et al 1997a). ............7.181
Figure 7.101: Idealised model of deformation during historically large drawdown..7.182
Figure 7.102: Walshaw Dean (Lower) dam, crest settlement versus time (Tedd et
al 1997a)..............................................................................................................7.183
Chapter 7: Deformation behaviour of embankment dams Page xiii

Figure 7.103: Cross section through Beliche dam at the main section (Maranha
das Neves et al 1994) ..........................................................................................7.189
Figure 7.104: Cross section of Chicoasen dam at the main section (Moreno and
Alberro 1982)......................................................................................................7.189
Figure 7.105: Section at Hirakud dam (Rao and Wadhwa 1958) ..............................7.192
Figure 7.106: Cumulative settlement during construction at Hirakud dam (adapted
from Rao 1957) ...................................................................................................7.192
Figure 7.107: Cross section of Nurek dam (adapted from Borovoi et al 1982).........7.193
Figure 7.108: Carsington dam, section at chainage 825 m after failure (adapted
from Skempton and Vaughan 1993). ..................................................................7.194
Figure 7.109: Carsington dam, vertical strain profiles from IVM during
construction, (a) IVM C at chainage 850 m, and (b) IVM B at chainage 705 m
(adapted from Rowe 1991)..................................................................................7.195
Figure 7.110: Carsington dam, variation of the maximum vertical strain in the
core with bank level (Rowe 1991). .....................................................................7.195
Figure 7.111: Maximum section at El Infiernillo dam (Marsal and Ramirez de
Arellano 1967). ...................................................................................................7.203
Figure 7.112: El Infiernillo dam, deformation instrumentation on station 0+135 at
the lower left abutment (Marsal and Ramirez de Arellano 1967).......................7.204
Figure 7.113: El Infiernillo dam, internal settlements during first filling at Station
0+135...................................................................................................................7.205
Figure 7.114: Main section at Chicoasen dam (Moreno and Alberro 1982)..............7.206
Figure 7.115: Chicoasen dam, inclinometer and cross-arm locations at the main
section (Moreno and Alberro 1982)....................................................................7.206
Figure 7.116: Internal settlement profiles during first filling at Chicoasen dam
(adapted from Moreno and Alberro 1982). .........................................................7.207
Figure 7.117: Beliche Dam, cross section (Maranha das Neves et al 1994)..............7.209
Figure 7.118: Beliche dam, post construction internal vertical settlement profile
within the embankment for the period from end of construction to 2.75 years
post construction. ................................................................................................7.209
Figure 7.119: Typical section of Canales dam (Bravo 1979). ...................................7.210
Figure 7.120: Canales dam, (a) cracking and differential settlement at the crest,
and (b) post construction settlement at the crest (Giron 1997) ...........................7.211
Figure 7.121: Main section at Djatiluhur dam (Sowers et al 1993)...........................7.212
Figure 7.122: Djatiluhur dam; (a) location of crack observed during construction,
January 1965; and (b) deformation of monuments at elevation 103 m, January
to April 1965 (Sherard 1973). .............................................................................7.214
Chapter 7: Deformation behaviour of embankment dams Page xiv

Figure 7.123: Eppalock dam, post construction settlement of SMPs on the crest
and slopes for the first five years after construction. ..........................................7.215
Figure 7.124: Main section at Blowering dam (courtesy of NSW Department of
Public Works and Services, Dams and Civil Section). .......................................7.218
Figure 7.125: Blowering dam, vertical strain during construction for selected
cross-arms intervals in the core...........................................................................7.220
Figure 7.126: Blowering dam, post construction internal settlement of the core
from IVM A during first filling. ..........................................................................7.221
Figure 7.127: Cross section of Ataturk dam (Cetin et al 2000) .................................7.221
Figure 7.128: Post construction settlement of the crest and downstream shoulder
at Ataturk dam. ....................................................................................................7.223
Figure 7.129: LG-2 dam; crack in crest and possible shear plane in core (Paré
1984) ...................................................................................................................7.224
Figure 7.130: Main section at Copeton dam (courtesy of New South Wales
Department of Land and Water Conservation) ...................................................7.225
Figure 7.131: Copeton dam, post construction internal settlement profile in the
core at IVM A. ....................................................................................................7.227
Figure 7.132: Copeton dam, post construction settlement between cross-arms
versus time. .........................................................................................................7.228
Figure 7.133: Eppalock dam, original design at maximum section (Woodward
Clyde 1999).........................................................................................................7.230
Figure 7.134: Eppalock dam, post construction settlement at main section and
CS1......................................................................................................................7.231
Figure 7.135: Eppalock dam, post construction horizontal displacement at the
main section. .......................................................................................................7.233
Figure 7.136: Djatiluhur dam, post construction crest settlement..............................7.234
Figure 7.137: Bellfield dam, main section at chainage 701 m (courtesy of
Wimmera Mallee Water).....................................................................................7.235
Figure 7.138: El Infiernillo dam, internal (a) settlement and (b) displacement in
the core from 1966 to 1972 (Marsal and Ramirez de Arellano 1972) ................7.236
Figure 7.139: El Infiernillo dam, post construction (a) settlement and (b)
displacement of surface markers.........................................................................7.237
Figure 7.140: Main section at Eildon dam (courtesy of Goulburn Murray Water). ..7.239
Figure 7.141: Eildon dam, post construction settlement of SMPs at chainage 685
m..........................................................................................................................7.240
Figure 7.142: Eildon dam, internal settlement profiles in core from IVM records
for the period 1981 to 1998.................................................................................7.240
Chapter 7: Deformation behaviour of embankment dams Page xv

Figure 7.143: Main section at Svartevann dam (Kjœrnsli et al 1982). ......................7.243


Figure 7.144: Svartevann dam, post construction (a) settlement and (b)
displacement. .......................................................................................................7.246
Figure 7.145: Cross section at Cougar dam (Cooke and Strassburger 1988). ...........7.247
Figure 7.146: Cougar dam; post construction (a) settlement and (b) displacement
normal to dam axis of the crest at the maximum section (adapted from Pope
1967). ..................................................................................................................7.248
Figure 7.147: Cougar dam; differential settlement and cracks on crest (Pope 1967) 7.249
Figure 7.148: Wyangala dam, post construction internal settlement in IVM B
located 9 m upstream of dam axis.......................................................................7.250
Figure 7.149: Section at Wyangala dam (courtesy of New South Wales
Department of Land and Water Conservation). ..................................................7.251
Figure 7.150: Rector Creek dam main section (Sherard 1953)..................................7.254
Figure 7.151: Rector Creek dam; post construction settlement and displacement
versus log time (adapted from Sherard et al 1963). ............................................7.254
Figure 7.152: Roxo dam, settlement versus log time (adapted from De Melo and
Direito 1982). ......................................................................................................7.255
Figure 7.153: Main section at Mita Hills dam (Legge 1970).....................................7.256
Figure 7.154: Main section at Dixon Canyon dam (courtesy of U.S. Bureau of
Reclamation). ......................................................................................................7.258
Figure 7.155: Post construction (a) settlement and (b) displacement of SMPs on
the upper upstream slope at the Horsetooth Reservoir embankments. ...............7.259
Figure 7.156: Belle Fourche dam; main section as constructed in 1911 (courtesy
of U.S. Bureau of Reclamation)..........................................................................7.260
Figure 7.157: Belle Fourche dam; post construction crest settlement over the
period 1911 to 1928 (to 17 years post construction)...........................................7.264
Figure 7.158: Belle Fourche dam closure section; post construction (a) settlement
and (b) displacement normal to dam axis of the crest and slopes over the
period 1985 to 2000 (73 to 89 years post construction). .....................................7.265
Figure 7.159: San Luis dam, section of the slide in the upstream slope in 1981 at
Station 135+00 (courtesy of United States Bureau of Reclamation). .................7.266
Figure 7.160: San Luis dam, difference between actual and predicted settlement
(Von Thun 1988).................................................................................................7.268
Figure 7.161: San Luis dam, comparison of settlement between SMPs at 13 m
upstream of dam axis and SMPs at 6.5 m downstream. ......................................7.268
Figure 7.162: San Luis dam, post construction settlement of SMPs in the vicinity
of the slide area. ..................................................................................................7.269
Chapter 7: Deformation behaviour of embankment dams Page xvi

Figure 7.163: Main section at Horsetooth dam (courtesy of the United States
Bureau of Reclamation). .....................................................................................7.271
Figure 7.164: Horsetooth dam, post construction displacement of SMPs near to
the main section...................................................................................................7.272
Figure 7.165: Pueblo dam; cross section of left abutment embankment at Station
75.........................................................................................................................7.273
Figure 7.166: Pueblo dam, post construction settlement of the left embankment
near Station 75.....................................................................................................7.274
Figure 7.167: Yan Yean dam; settlement rate versus time of SMPs at Chainage
150 and 750 m. ....................................................................................................7.279
Chapter 7: Deformation behaviour of embankment dams Page 7.1

7.0 THE DEFORMATION BEHAVIOUR OF


EMBANKMENT DAMS

7.1 INTRODUCTION TO THIS CHAPTER


Failures by slope instability constitute about 6% of failures and 37% of incidents of
embankment dams (Foster 1999). A number of these incidents were prevented from
becoming failures by early detection of the impending failure from visual inspection or
from deformation monitoring, and remedial action is taken by lowering the reservoir
and/or construction of stabilization works.
Those who are responsible for interpreting the monitoring data do not have clear
criteria to assess their data and have to rely on their judgement and personal experience.
Currently available numerical modelling packages and the sophistication of the
constitutive models on which they are based are improving making for useful tools in
modelling embankment dam behaviour. However, few reported Type A predictions
have been able to accurately model the deformation behaviour during construction and
on first filling. Methods based on comparison to historical records of similar
embankments are still heavily relied upon and, particularly for the assessment of post
construction deformation behaviour, provide the responsible personnel with the best
guide to assessment of embankment performance.
Overall, the purpose of the presented data is, from a broad database of well-
instrumented earth and earth-rockfill embankments, to provide dam owners and their
consultants with methods that:
• Allow comparison of their structures to similar embankment types in terms of the
deformation behaviour during construction, on first filling and post first filling.
• Broadly define “normal” deformation behaviour and from this platform to then
identify potentially “abnormal” deformation behaviour, either in terms of
magnitude, rate or trend.
• Provide some guidance on the trends in deformation behaviour that are potentially
indicative of a marginally stable to unstable slope condition, and precursors to slope
instability.

The embankment types considered in this chapter include earthfill, zoned earthfill,
zoned earth and rockfill (mainly central core earth and rockfill) and puddle core earthfill
Chapter 7: Deformation behaviour of embankment dams Page 7.2

embankments. A brief summary of the literature related to slope instability statistics


and prediction of deformation behaviour is presented in Section 7.2. The database of
case studies is summarised in Section 7.3. Sections 7.4 to 7.7 present the general
deformation behaviour for the embankment types considered, and Sections 7.8 to 7.12
the methods for identification of “abnormal” deformation behaviour with reference to
case studies.
Section 7.13 presents a summary of the methods of identification of “abnormal”
deformation behaviour and highlights important aspects evident from the case study
analysis. Guidelines are also presented on trends in the deformation behaviour that are
potentially indicative of a marginal stability condition. Section 7.14 presents a
summary of the outcomes from the analysis for use in prediction of, or comparison of
the deformation behaviour of an embankment during and post construction.

7.2 LITERATURE R EVIEW

7.2.1 FAILURE AND ACCIDENT STATISTICS

This review of slope instability incidents in embankment dams is a summary of that


presented in Section 5.2.1 of Chapter 5 as part of the post-failure deformation analysis
of instability incidents in embankment dams. Most of the data and findings on slope
stability incidents is a summary of the work by Foster (1999) and Foster et al (2000) on
dam failure and accident incidents, derived from a database of some 11192 large
embankment dams from around the world constructed up to 1986.
Foster (1999) reported 124 incidents of slope instability. Of these, 12 were failures
that involved a breach and uncontrolled release of water from the reservoir and the
remaining 112 were accidents (no breach). They concluded that the major factors
influencing the likelihood of slope instability in embankment dams were:
• Slope instability is more likely in earthfill embankments than earth and rockfill
embankments. Homogeneous dam embankments (without filters) and hydraulic fill
dam embankments are statistically the most susceptible to slope instability.
• Slope instability is more likely for dam embankments on soil foundations than for
dam embankments on rock foundations. Of the slides passing through the
embankment and foundation, the incidence of slope instability is much greater in
foundations of high plasticity clays.
Chapter 7: Deformation behaviour of embankment dams Page 7.3

• Slope instability is much more likely in dam embankments constructed of high


plasticity clays and silts, particularly for slides in the downstream slope. Dam
embankments constructed of sands and gravels have a significantly lower incidence
of slope instability.
• Compaction, or lack there of, of the earthfill materials is also a significant factor.
The incidence of slope instability is much more likely for earthfills constructed by
placement with no formal compaction, and the least likely for rolled, well
compacted earthfills.

Further analysis as part of this research project (refer Section 5.2.1) found that:
• Earthfill embankments (homogeneous earthfill, earthfill with filters and earthfill
with rock toe) made up a very high percentage (80%) of the slope stability incidents
that passed through the embankment only.
• For embankments that incorporate rockfill zones, it is very to extremely unlikely
that the surface of rupture will pass through the rockfill zone itself, it will
preferentially pass through the foundation.

7.2.2 HISTORICAL DEVELOPMENT OF EMBANKMENT DAMS AS THIS AFFECTS


DEFORMATION BEHAVIOUR

7.2.2.1 Earth and Earth-Rockfill Dams

An historical summary of the use of rockfill in embankment design and construction


(Galloway 1939; Cooke 1984; Cooke 1993) is given in Table 6.2 in Section 6.3.1 (of
Chapter 6).
For earthfills, the significant changes affecting deformation behaviour were
developments in placement methods. The use of rollers from about the 1850’s was
significant, although their usage was more widespread from about the late 1890’s into
the early 20th century. But, the most significant development was after the 1930’s with
the application of soil mechanics principles (Wilson and Squier 1969; Penman 1986)
and consideration of moisture content control, compacted layer thickness and level of
compaction. Consideration of these principles was evident in embankment construction
earlier in the 20th century (Schuyler 1912), such as at Belle Fourche dam in California.
Chapter 7: Deformation behaviour of embankment dams Page 7.4

7.2.2.2 Puddle Core Earthfill Dams


Skempton (1990) referred to the puddle core earthfill dam as the preferred method of
dam construction in the United Kingdom (UK) in the 19th and first half of the 20th
century. It was also the preferred method of construction in India and Australia up to
the 1930’s.
The significant historical events in the design and construction of puddle dams in
the UK (Skempton 1990; Foster 1999) are summarised as follows:
• The first puddle dams were constructed in the late 18th century.
• Puddle Core - improvements in preparation procedures resulting in greater
uniformity of the core material
− Late 18th century to mid 19th century the puddle core was watered on the
embankment, allowed to soak and then heeled (by foot) or rammed (using hand
held rammers) in 150 to 200 mm layers.
− In the second half of the 19th century the clays for core material were generally
wetted up in the borrow area before being brought onto the embankment.
− 20th century, preparation of core materials in pugmills and placement in 100 to
150 mm layers.
• Shoulder Fill - significant improvements in design and method of placement
− Initially, no zoning in shoulder fill and no formal compaction. Compaction was
typically using loaded wagons and carts in relatively thick layers (generally less
than 1 m but up to 1.5 - 2 m).
− From the 1850’s/60’s, following the failures of Bilberry Dam in 1852 (by
internal erosion) and Dale Dyke in 1864 (by a combination of internal erosion
and over-topping), embankment designs generally included a select zone of the
more cohesive earthfills adjacent to the core.
− From the 1900’s improved compaction of shoulder fill using thinner layers and
steam driven rollers.
− From the 1940’s further improvements in compaction procedures following
improvements in earth-moving plant and the application of soil mechanics.
• The use of rolled clay cores took over from puddle clay cores in the UK in the
1960’s.
Chapter 7: Deformation behaviour of embankment dams Page 7.5

In Australia and India the use of rolled clay cores took over from puddle clay cores
in the 1930’s. The last large puddle dam constructed in Australia was in the 1930’s.
Improvements in the preparation of puddle cores and alteration in design and placement
methods of the shoulder fill occurred in Australia at relatively similar time periods to
those in the UK. There is evidence of the use of horse or bullock drawn sheepsfoot
rollers (up to 3.5 tonne dead-weight) for compaction of the shoulders and bullocks for
compaction of the puddle core in Australia from the 1870’s.

7.2.3 FACTORS AFFECTING THE DEFORMATION OF EMBANKMENT DAMS

7.2.3.1 Deformation Properties of Rockfill

An assessment of the factors affecting the compressibility of rockfill is presented in


Section 6.3.2 of Chapter 6. In summary, field observations (Mori and Pinto 1988; and
others) and the results of large scale laboratory tests (Marsal 1973; Marachi et al 1969;
and others) indicate the compressibility properties of rockfill are affected by:
• Degree of compaction of the rockfill;
• Applied stress conditions and stress path;
• Particle shape and particle size distribution; and
• Intact strength of the rock used as rockfill.

An important aspect in the evaluation of the deformation behaviour of rockfill in


embankment dams is its susceptibility to collapse compression on wetting. This is
discussed in Sections 7.5.2 and 7.10.1.
The time dependent or creep type deformation of rockfill is an important aspect for
evaluation of the post construction deformation behaviour in embankment dams.
Guidelines based on case study analysis of field records of concrete face rockfill dams
are given in Sections 6.4.4 and 6.5.3.

7.2.3.2 Deformation Properties of Earthfill

The general features of the stress-strain relationship of soils are well described in
textbooks on soil mechanics and are applicable to describe the deformation behaviour of
rolled earthfills. Factors such as the degree of over-consolidation, permeability, soil
structure, matric suction and the rate, magnitude and direction of loading affect the
Chapter 7: Deformation behaviour of embankment dams Page 7.6

strength and compressibility properties of an earthfill. Some of these factors are


affected by:
• The soil type, including its mineralogy, gradation and plasticity;
• The degree of compaction; and
• The moisture content at placement relative to Standard optimum.

Gould (1953,1954) found that a reasonable prediction of the drained compressibility


of a well compacted, rolled earthfill could be estimated from its classification and
plasticity of the fines fraction. For more plastic soil types, he found that the moisture
content relative to the Standard Proctor optimum influenced the shape and magnitude of
the compressibility curve. Most of Gould’s data was of earthfills placed on the dry side
of Standard Proctor optimum.
The compacted earthfill is a partially saturated soil with an amount of over-
consolidation that is related to the compactive effort of the roller and the pore water
suction. The pore water suction is dependent on the moisture content at placement, the
degree of saturation of the compacted earthfill and the material type. For a dry placed,
well-compacted earthfill, the amount of settlement increases with the increase in
effective stress as the embankment is constructed. The settlement occurs due to
compression of pore air in the partially saturated soil and is a relatively rapid process, as
demonstrated by the fact that most of the earthfill’s settlement occurs during
construction. As the pore air is compressed the degree of saturation increases and pore
water suction decreases. At some point, positive pore water pressures may be
developed. Laboratory data presented by Fredlund and Rahardjo (1993) and Holtz and
Kovacs (1981) indicate that the value of the pore water pressure parameter, B, is low for
degree of saturation up to 85 to 90% and possibly higher.
In compacted earthfills placed close to or wet of Standard optimum moisture
content, the degree of saturation is relatively high at compaction (generally greater than
85 to 90%) and the matric suction relatively low compared with dry placed earthfills.
Positive pore water pressures are generally developed at low levels of confining stress
and the pore water pressure parameter, B, approaches high values, from about 0.6 to
almost 1.0. Assuming the earthfill to be of low permeability, conditions during
conditions will be close to undrained. For these earthfill types deformations largely
occur as plastic type deformations due to yielding in undrained loading.
Chapter 7: Deformation behaviour of embankment dams Page 7.7

On filling, seepage through the earthfill begins. It may take many years to reach a
steady seepage condition, depending on the permeability of the partially saturated
and/or saturated earthfill (refer Section 7.6.3, item f). As the degree of saturation of the
partially saturated earthfill increases on saturation, deformations will occur.
Deformations due to collapse compression on wetting can occur if the rolled earthfill is
placed well dry of Standard optimum (even if heavily compacted) or dry placed and
poorly compacted (refer Sections 7.5.2 and 7.13). In the long term, there will be a small
time dependent creep component to the deformation behaviour. This is generally
greater for poorly compacted soils than for well-compacted soils.

7.2.3.3 Summary of Research on Puddle Core Earthfill Dams


Building Research Establishment (BRE) and Imperial College have, since the 1980s,
published the findings of research into the performance of puddle core earthfill dams.
In the area of deformation behaviour the research has been mainly targeted at the long-
term performance of these embankments more than 50 to 100 years after construction.
Instrumentation installed in a number of puddle core embankments from about the late
1970’s/early 1980’s comprised monitoring of:
• Pore water pressures, horizontal stresses and internal deformations (both vertical
and horizontal) in the puddle core and in some cases in the puddle clay cut-off.
• Pore water pressures in the downstream fill.
• Surface deformations of the crest, upper part of the upstream slope and downstream
slope.

A set of puddle dams owned by Yorkshire Water, including Ramsden, Walshaw


Dean Lower, Ogden and Holmestyes dams, have been reasonably well monitored and
analysed by BRE and Imperial College in conjunction with the owners. Imperial
College has undertaken finite element analysis of these and other puddle core earthfill
embankments (Tedd et al 1997b; Dounias et al 1996; Kovacevic et al 1997).
Table 7.1 summarises the mechanisms identified by Tedd et al (1997a) affecting the
long-term deformation behaviour of puddle core earthfill embankment dams. They
comment that an understanding of the likely deformation behaviour due to processes not
detrimental to dam safety allows for easier identification of deformations that may be
indicative of a “hazardous” situation.
Chapter 7: Deformation behaviour of embankment dams Page 7.8

Tedd et al (1997a) defined the factors affecting the magnitude and/or direction of
“normal” long-term deformation behaviour of puddle core earthfill dams as:
• The age of the dam;
• The height of the dam;
• The position of the element controlling the phreatic surface within the dam. For
most puddle dams the central puddle core itself is the element controlling the
phreatic surface within the dam, but there are several embankments where an
upstream lining is the controlling element (e.g. Holmestyes Dam).
• The compressibility properties of the various embankment zones and foundation;
• The permeability of the earthfill/s in the upstream shoulder; and
• The reservoir operation, in particular the past history of operation, the reservoir level
at the time of measurement and the duration and depth of drawdown.

Table 7.1: Mechanisms affecting the long-term deformation behaviour of old puddle
core earthfill embankments (Tedd et al 1997a).
Mechanisms Causing “Normal” Deformation Deleterious Mechanisms /
Behaviour Processes
Shrink-swell related deformations due to seasonal Slope instability
fluctuations in moisture content. Internal erosion processes
Secondary consolidation of the core and creep of the
shoulder fill under “steady state” conditions.
Deformations due to stress changes associated with the
fluctuation in reservoir level during normal operation.
Consolidation of the foundation.

The case study data and some of the findings from the published literature by BRE and
Imperial College relating to the deformation behaviour of puddle core earthfill dams is
included with the analysis and discussion in this chapter (refer Sections 7.7, 7.12, 7.13
and 7.14.4). Some of the findings of the research by BRE, other than from the
monitored field deformation behaviour, are:
• Laboratory oedometer tests on samples of the core and upstream fill (Holton 1992;
Tedd et al 1994; Tedd et al 1997b) showed:
− Measured cα, coefficient of secondary compression, were in the range 0.0001 to
0.0032. The general trend was for cα to increase with increasing effective
vertical stress.
Chapter 7: Deformation behaviour of embankment dams Page 7.9

− Non-linear stress-strain behaviour on repeated unloading and reloading on


remoulded samples (at stress levels below the initially applied stress). The
stress-strain behaviour followed hysteresis loops, with the constrained modulus
value decreasing as the number of reload cycles increased, and each load cycle
resulted in a non-recovered strain, the magnitude of which was dependent on the
magnitude of the load.
• Where the shoulder fill has been sourced from weathered sedimentary deposits it is
relatively permeable, although the permeability is quite variable (Tedd et al 1993).
These fills generally comprised a mixture of silty and sandy clays with varying
proportions of gravel to boulder size fragments of mudstone, siltstone and
sandstone. Tedd and Holton (1987) report the permeability from constant head tests
to vary from 2 × 10 -5 m/s to 10 -7 m/s in the downstream shoulder fill at Ramsden
Dam. For Cwnwernderi Dam, Charles and Watt (1987) reported a relatively rapid
response in horizontal earth pressure on reservoir filling indicating the upstream
filling to be quite permeable.
• Charles and Watt (1987) found that the horizontal stresses in the core of puddle
dams some 100 years post construction were relatively low, significantly lower than
nominal stresses for a homogeneous type embankment, and Charles and Tedd
(1991) comment that measured vertical stresses were not significantly greater than
the horizontal stresses. These findings confirm the low stresses in the core present
at the end of construction due to arching are still present some 100 years after
construction.

7.2.4 PREDICTIVE METHODS OF DEFORMATION BEHAVIOUR

For embankment dams, the components of deformation are those that occur as a result
of changes in effective stress conditions (during construction, on impoundment and due
to reservoir fluctuation), changes in total stress (for wet placed earthfill cores in zoned
embankments), on saturation or wetting (e.g. collapse compression) and the on-going
time dependent or creep type deformations. Predictive methods are typically divided
into the three components of deformation during construction, on first filling and long-
term post construction (or post first filling).
Most predictive methods cover the deformation behaviour of one or two of these
components. Finite element analyses have been used (Saboya and Byrne 1993;
Chapter 7: Deformation behaviour of embankment dams Page 7.10

Kovacevic 1994; Naylor et al 1997; Dounias et al 1996, amongst others) for analysis of
deformation due to changes in stress conditions, mostly for deformation during
construction and first filling, but also for deformation under large drawdown. Empirical
predictive methods, usually based on historical performance of embankments, are more
generally available for post construction deformation some of which incorporate the
deformation on first filling, although methods are available for deformation during
construction.

7.2.4.1 Empirical Predictive Methods


(a) Predictive Methods of Deformation During Construction
There are methods available for prediction of deformations during construction, but they
are limited in number. Development of these methods has probably been curtailed by
the availability of finite element methods in the last 30 to 40 years, their relative ease of
use to model embankment construction and improvements in constitutive relationships
to model the material compressibility and deformation behaviour. In addition, the
complexity of stress and deformation behaviour during construction, particularly for
zoned embankments, makes assessment by simple empirical methods difficult. Several
of the methods developed are:
• Poulos et al (1972) developed charts for estimation of stresses and vertical and
horizontal deformations during construction based on elastic solutions for a
homogeneous embankment on a rigid foundation. The methods provide a quick and
simple means of estimating deformations, which the authors comment are suitable
as a preliminary estimate.
• Penman et al (1971) developed a method for estimation of the settlement during
construction under the dam axis based on elastic solutions, but that allows for
material stress-strain non-linearity by using an equivalent compressibility. It is
based on the assumption of confined conditions (i.e. zero lateral strain) under the
dam axis and is therefore suitable where this assumption is reasonable (e.g.
homogeneous embankment).
• ICOLD (1993) derived a general equation for the vertical strain of rockfill in terms
of vertical stress, rockfill modulus and creep strain parameters (Equation 7.1) based
on the assumption of a linear relationship between stress and strain of coarse rockfill
and non-linear relationship between strain and time. Derivations are given (refer
Chapter 7: Deformation behaviour of embankment dams Page 7.11

ICOLD 1993) for the settlement at a specific elevation and period of time after start
of construction, settlement at end of construction and settlement post construction.
• Several other methods have been developed for prediction of the lateral deformation
of the shoulder during construction, including:
− Resendiz and Romo (1972) and Walker and Duncan (1984) for earthfill or
essentially homogenous dams.
− Penman (1986) provides guidelines for assessment of “acceptable” and
“excessive” rates of horizontal displacement, mainly as guide for evaluation of a
possible impending failure condition. It is based on a limited number of case
studies of differing embankment type and should be used with caution.

σ t
e= +σ (7.1)
EM θ +λt

where e = relative strain, σ = vertical stress (in MPa), EM = modulus of instantaneous


strain of the rockfill (in MPa), t = time (in years), and θ and λ are the empirical
parameters describing creep strain (in MPa/year and MPa respectively).
One of the major issues with deformation during construction relates to material
compressibility properties under the imposed stresses from the stress path during
construction. The stress path and stress conditions from field instrumentation (Charles
1976) and finite element analysis (Saboya and Byrne 1993; Kovacevic 1994; Charles
1976; and implicit in the analysis by others) are shown to vary depending on the
embankment zoning geometry, material properties and location within the embankment.
Hence, the dependency on finite element analyses to model the stress path, stress
conditions and deformation behaviour in all but the simplest problems. The assumption
of confined conditions under the central embankment region for homogeneous
embankments or embankments with broad central zones is shown by field
instrumentation and finite element analysis to be a reasonable assumption. For these
conditions published data is available for estimation of material moduli of earthfills
from Gould (1953, 1954).
Gould (1953, 1954) analysed the compressibility characteristics of rolled earthfill
from the internal deformation records of some 22 rolled earthfill and earth-rockfill
embankments. All case studies were USBR embankments and predominantly of
embankments with thick to very broad and well compacted rolled earthfill cores. Gould
Chapter 7: Deformation behaviour of embankment dams Page 7.12

found that the classification of the earthfill and plasticity of the fines fraction provided a
reasonable prediction of compressibility. For the more plastic soil types, he found
moisture content relative to the Standard Proctor optimum influenced the shape and
magnitude of the compressibility curve. Table 7.2 presents a summary of Gould’s data
(adapted from Sherard et al 1963).
Giudici et al (2000) and Pinto and Marques Filho (1998) provide methods for
estimation of the rockfill moduli during construction from the deformation behaviour of
concrete face rockfill dams. These methods are summarised in Section 6.3.3.2 of
Chapter 6. What is considered an improved method for prediction of rockfill moduli is
presented in Sections 6.4.2 and 6.5.1.

(b) Predictive Methods of Deformation Behaviour Post Construction


Empirical predictive methods of the post-construction deformation from the literature
are typically based on historical deformation curves for similar embankment types or
empirical relationships derived from historical deformation performance. Several of the
methods incorporate the deformation during first filling whilst others only consider the
prediction of deformation post first filling.

Table 7.2: Vertical compression of rolled, well-compacted earthfills measured during


construction (adapted from Sherard et al 1963)
Approximate Range of Measured
Vertical Compression (%) *1
Embankment Soil Type
At 10 psi (approx. At 100 psi (approx.
70 kPa) *2 700 kPa) *2
Silty gravels and coarse silty sands (GM and SM). 0.2 to 0.3 0.9 to 1.4
Fine silty sands and silts of low plasticity (SM/ML). 0.2 to 0.5 1.2 to 2.1
Clayey sands and gravels (GC and SC). 0.3 to 0.8 *3 1.9 to 3.3
Clays of low to medium plasticity (CL and CL/ML). 0.2 to 1.1 *3 2.8 to 4.6 *4
*1 vertical compression measured from internal settlement gauges within the embankment
*2 pressures are vertical effective stresses
*3 variation directly with placement moisture content relative to Standard Proctor optimum, low value for
soils placed well dry of optimum and high value for soils placed wet of optimum (Gould 1954).
4
* spread influenced by moisture content at placement and plasticity of the fines fraction (Gould 1954).

Table 7.3 provides a summary the range of post construction deformations of earthfill
and earth and rockfill dams from the published literature. The data indicates that the
Chapter 7: Deformation behaviour of embankment dams Page 7.13

post construction settlements and displacements are generally relatively small for well-
compacted earthfills and rockfills, and are much greater for embankments incorporating
dumped rockfill.
Sowers et al (1965), Soydemir and Kjærnsli (1975), Clements (1984), Dascal
(1987), ICOLD (1993) and Charles (1986), amongst others, have proposed predictive
methods of post construction deformation behaviour. All have been derived from the
historical performance of a database of embankments, in most cases from dams
predominantly of rockfill (concrete or membrane face, and sloping and central core
earth and rockfill dams). The method proposed by Charles (1986) is for puddle core
earthfill embankments.
The methods by Clements (1984), Dascal (1987) and the data summarised in Table
7.3, provide a range or bounds of likely deformation behaviour, with Clements having
the largest database of case studies. Both Dascal and Clements assume zero time is the
time of the first deformation reading after construction. Clements (1984) suggested the
bounds be used to provide an expected range of deformation, and for better prediction
he suggested estimating from the deformation curves of dams with similar
characteristics. Pinto and Marques Filho (1985) commented that these methods are
useful, however, their application is not so straightforward as the deformation behaviour
is strongly dependent on the initial time of measurement in relation to the end of
construction. And, as pointed out by Parkin (1977), the shape of the deformation versus
time curve will vary depending on the established zero time.
The other methods have developed empirical relationships for estimation of
settlement and/or displacement. The methods by Sowers et al (1965) and Soydemir and
Kjærnsli (1979), and comments by Parkin (1977) on strain rate analysis are summarised
in Section 6.3.3.2 of Chapter 6. The Russian method referenced by ICOLD (1993) and
Charles (1986) are summarised below. Most of the proposed equations are simple
logarithmic or power type relationships based on one or two variables.
The Russian method for estimation of settlement (referenced in ICOLD (1993)),
which is based on time after the end of construction, stress level and the creep strain
empirical parameters, is quite complex in comparison to other proposed relationships.
Equation 7.1 gives the basic form of the equation. ICOLD (1993) provides further
details on this method.
Chapter 7: Deformation behaviour of embankment dams Page 7.14

Table 7.3: Published ranges of post construction deformation of embankment dams


Reference Dam Type/s Deformation Parameter Range of Deformation Comments
(% of dam height) *1
ICOLD (1993) “rockfill” * 2 crest settlement 0.2 to 1.0% Crest displacement up to 50% of the
crest displacement 0.1 to 0.5% crest settlement.
shoulder settlement 0.1 to 0.2%
Sowers et al (1965) “rockfill” * 2 crest settlement 0.25 to 1.0% 14 dams, settlements up to 10 years
(upper range for dumped RF) after construction.
Clements (1984) CFRD crest settlement Up to 2.5% (dumped RF) Database of 68 dams.
0 to 0.25% (compacted RF)
crest displacement 0 to 2.5%
CCER crest settlement 0.05 to 1.25%
crest displacement -0.75% to 0.5%
Sloping core crest settlement 0.06 to 1.1%
crest displacement 0 to 0.6%
Bernell (1958) CCER with moraine crest settlement (during first filling) 0 to 0.2% (silty and sandy moraines) 6 dams with moraine cores placed
core 0.05 to 0.3% (clayey moraines) using the wet compaction method.
Fine fractions less than 20%.
Dascal (1987) “rockfill” dams (*2) crest settlement < 0.35% (compacted RF) 15 Hydro Quebec dams and dikes on
with moraine cores 0.3 to 0.55% (dumped RF) rock foundations
crest displacement ≥ crest settlement for compacted RF
< crest settlement for dumped RF
downstream shoulder settlement up to 0.7 to 0.8%
Sherard et al (1963) “rockfill” * 2 crest settlement 0.1 to 0.4% (for well constructed wetted RF) Greater settlement for dumped RF.
well constructed dams crest displacement on first filling < 25 to 50 mm Greater displacements for dams with
dumped RF.
Gould (1954) rolled earthfill dams crest settlement < 0.2% in first 3 years Typical range of settlement for USBR
<0.4% up to 14 years dams.
*1 displacement is horizontal deformation, downstream displacement is positive and upstream is negative.
*2 “rockfill” dams including membrane face rockfill dams, and central and sloping earth core rockfill dams
CFRD = concrete face rockfill dam CCER = central core earth and rockfill dam RF = rockfill USBR = United States Bureau of Reclamation
Chapter 7: Deformation behaviour of embankment dams Page 7.15

Charles (1986) proposed a settlement index, SI, for assessment of the long-term crest
settlement behaviour of puddle core earthfill embankments (Equation 7.2). Charles and
Tedd (1991) point out that the settlement index is analogous to the coefficient of
secondary consolidation for a clay soil. They provide values of the settlement index for
a number of puddle core embankments during the period of normal reservoir operation.
Tedd et al (1997b) suggest that values of settlement index “greater than 0.02 could
indicate some mechanism other than creep was causing the settlement and that the
situation should be seriously examined”.

s
SI = (7.2)
1000 ⋅ H ⋅ log (t 2 t1 )

where s is the crest settlement in millimetres measured between times t2 and t1 after the
completion of embankment construction, and H is the height of the dam in metres. The
equation is of the same form as that proposed by Sowers et al (1965).

7.2.4.2 Numerical Methods


Duncan (1996) and Kovacevic (1994) are recent references on the state of the art in
finite element analyses applicable to the deformation behaviour of embankments,
mainly zoned earth and rockfill dams. They discuss the methods of analysis, their
limitations, available constitutive models of the stress-strain relationship and areas of
uncertainty.
Duncan (1996) comments that the choice of stress-strain constitutive model used in
the analysis is a balance between simplicity and accuracy, and that the choice of
constitutive model will depend on the purpose of the modelling. He commented that if
the purpose of the analysis is to analyse stresses and/or the trend of deformations then
more simplistic models could be suitable, depending on the relative stiffness between
different embankment materials. More accurate modelling of deformations requires a
more complex stress-strain model that more realistically approximates the real
behaviour of the rockfill and earthfill.
Aspects of finite element analysis (after Duncan (1996) and Kovacevic (1994)) as
they relate to modelling rockfill are summarised in Section 6.3.3.1 of Chapter 6.
Important aspects raised by Duncan (1996) and Kovacevic (1994) related to modelling
in general and for earthfills more specifically are:
Chapter 7: Deformation behaviour of embankment dams Page 7.16

• The importance of modelling the construction as a series incremental layers


• The type of stress-strain relationship used for modelling materials. Duncan (1996)
comments that elastic stress-strain models (e.g. linear elastic, multi-linear elastic or
hyperbolic models) are appropriate where the soils are not stressed to failure, but,
that these models are not suitable for modelling undrained behaviour or problems
where local failure occurs. He recommends the use of models that incorporate
plasticity theory for these situations (e.g. for wet placed earthfill cores of thin to
medium width in zoned embankments).
• For situations where the earthfills are not stressed to failure, linear elastic models are
generally not suitable for accurate modelling of deformation behaviour due to the
non-linear stress-strain relationship of earthfill and rockfill. For these situations, the
use hyperbolic or multi-linear elastic models provide improved predictions of
deformation and have been successfully used, particularly for deformation
prediction during embankment construction.
• A limitation of hyperbolic and multi-linear elastic models is that the model
parameters are typically derived from triaxial compression or oedometer laboratory
tests and are therefore limited by the narrow range of stress paths covered in
comparison to the broader range of stress paths imposed in the field (Kovacevic
1994).

An important component of the modelling of embankment dams is the consideration


of collapse compression of susceptible rockfills and earthfills on wetting. This is
discussed further in Section 6.3.3.1 of Chapter 6.
The effect of pore water pressure dissipation in earthfills during construction has
been considered by a number of authors including Eisenstein and Law (1977), in
modelling Mica dam, and by Cavounidis and Höeg (1977), amongst others. For these
cases, the incremental embankment construction was modelled as a two stage process,
the first stage modelling the new layer construction using undrained properties for the
core and the second stage modelling pore water pressure dissipation.
In most cases though, pore water pressure development in the core is ignored. For
wet placed earthfills, where high pore water pressures are developed during
construction, the core is often modelled using undrained strength and compressibility
parameters, and the permeability is assumed as sufficiently low such that pore water
pressures will not dissipate during the period of construction. For dry placed earthfills,
Chapter 7: Deformation behaviour of embankment dams Page 7.17

the earthfills are often modelled using elastic models (e.g. multi-linear elastic or
hyperbolic elastic). In some cases, the earthfill is analysed in terms of effective stresses
and using elasto-plastic models with pore water pressures input into the model; e.g.
Naylor et al (1997) for the core of Beliche dam. In this case, the analysis was
undertaken after the embankment had been constructed and actual pore water pressures
were used to derive the input to the model.

7.3 DATABASE OF CASE S TUDIES


The case study database comprises some 134 embankments; 63 central core earth and
rockfill dams, 21 zoned earth and rockfill dams, 23 zoned earthfill dams, 10 earthfill
dams and 17 puddle core earthfill embankments. A summary of the case studies in the
database is presented in Table 7.4 to Table 7.8.

7.3.1 EARTHFILL AND ZONED EARTH AND EARTH-ROCKFILL EMBANKMENTS

The type of rolled earthfill and zoned earth and earth-rockfill embankments (Figure 1.6
in Section 1.3.3 of Chapter 1) in the case study database include:
• Central core earth and rockfill embankments (Table 7.4), this is the dominant
embankment type;
• Zoned earth and rockfill embankments (Table 7.5);
• Zoned earthfill embankments (Table 7.6); and
• Earthfill embankments (Table 7.7), including homogeneous earthfill, earthfill with
filters and earthfill with rock toe.

Embankments with unusual design or of design with limited numbers of case studies
available from the literature were excluded. Sloping core zoned earth and rockfill
embankments have not been considered.
Further details on most of the case studies are given in the tables in Section 1 of
Appendix F, including a summary of the embankment details, material types and their
method of placement, reservoir operation, hydrogeology, monitoring (during and post
construction) and references for each dam. In most embankments the main earthfill
zone acting as the water barrier has generally been placed in thin layers and well
compacted using rollers suitable for the earthfill type.
Chapter 7: Deformation behaviour of embankment dams Page 7.18

Approximately half of the earthfill and zoned earth and earth-rockfill case studies
are embankments owned or supervised by sponsors and in-kind sponsors of the research
project. For most of these embankments the deformation records are of good quality.
These case studies have been supplemented by case studies from the published literature
for which the deformation records were of reasonably to good quality, and a reasonable
level of detail was available on the embankment design, material types and construction
methods.

7.3.2 PUDDLE CORE EARTHFILL EMBANKMENTS

A database of seventeen case studies of puddle core earthfill dams (3 from Australian
and 14 from the UK) has been gathered for analysis of deformation behaviour (Table
7.8). Further details on the case studies are given in Section 1 of Appendix F, including
a summary of the embankment details, reservoir operation, hydrogeology and
monitoring for each dam.
Most of the available deformation records are from surface measurement points
(SMP) established many years after construction, in the 1970’s to 1980’s for a number
of the case studies. Some monitoring data and anecdotal information is available on the
early performance of several embankments. Compared to the rolled earthfill and zoned
earth and earth-rockfill embankment case studies, the monitoring records for the puddle
core earthfill dam data set is fairly limited, but to be expected given the age of most of
the dams.

7.4 GENERAL DEFORMATION B EHAVIOUR DURING CONSTRUCTION OF


EARTHFILL AND ZONED EARTH AND EARTH -ROCKFILL
EMBANKMENTS
For virtually all of the embankments considered the main earthfill zone/s have generally
been placed in thin layers and well compacted in line with modern practice. In some of
the older embankments the moisture content control of the earthfill was possibly not as
stringent as it would have been with modern practice, but this has not necessarily been
detrimental to the embankment performance.
Chapter 7: Deformation behaviour of embankment dams Page 7.19

Table 7.4: Central core earth and rockfill embankments in the database
n
Height Embankment Class Rockfill
1
Dam Name Country * H Core Core
n
(m) Class Source Compaction Rating
Size Type
Agigawa Japan 102 5,1,1 c-tm CL - good (?)
Ajuare Sweden 46 5,2,2 c-tn SM schists & gneisses poor
Ataturk Turkey 184 5,2,0 c-tm CH basalt & limestone reasonable (?)
Bath County Upper
USA, Virginia 134 5,2,0 c-tm SC (?) siltstone & sandstone good (?)
Dam

Beliche Portugal 55 5,2,2 c-tn GC schists & greywacke poor (?)

Bellfield Australia, Vic. 40 5,2,0 c-tm CL/SC sandstone poor, dry placed
Bjelki-Peterson Australia, Qld 41.5 5,2,1 c-tn CL phyllite good
Blowering Australia, NSW 112 5,2,0 c-tm SC - CL quartzite & phyllite good
Chaffey Australia, NSW 54 5,2,0 c-tm CH/GC jasper reasonable to good

Cherry Valley USA, San Francisco 100 5,2,0 c-tk SM/ML granite & granodiorite dumped and sluiced

Chicoasen Mexico 261 5,2,0 c-tn GC limestone good

Copeton Australia, NSW 113 5,2,0 c-tm SC granite reasonable to good, dry placed

quartzite, sandstone &


Corin Australia, ACT 74 5,2,0 c-tm SM good
siltstone
good (downstream), reasonable
Cougar USA, Oregon 159 5,2,0 c-tn (u) GM basalt and andesite
(upstream)
Dalesice Checkoslovakia 90 5,2,0 c-tn (u) CL - good
Dartmouth Australia, Vic. 180 5,2,0 c-tm SC/SM granitic gneiss good, dry placed
Djatiluhur Indonesia 105 5,2,0 c-tm (u) CH andesite poor
Eildon Australia, Vic. 79 5,1,0 c-tk CL quartzitic sandstone poor, dry placed (?)
diorite & silicified
El Infiernillo Mexico 148 5,2,0 c-tn CL-CH poor to reasonable, dry placed
conglomerate
Eppalock Australia, Vic. 47 5,2,0 c-tk CL basalt poor, dry placed
Frauenau Germany 80 5,2,0 c-tn SM - -
Fukada Japan 55.5 4,1,2 c-tm GC (?) sandstone & granite. good (?)
reasonable to good (inner),
Geehi Australia, NSW 91 5,2,0 c-tn SM granitic gneiss
poor (outer)
Gepatsch Austria 153 5,2,0 c-tn GM/GC gneiss reasonable, dry placed

Glenbawn Saddle A Australia, NSW 35 5,2,0 c-tm - limestone & sandstone good (?)

Glenbawn - main
dam (original, prior to Australia, NSW 76.5 5,1,0 c-tk CL limestone poor, dry placed
raising)
Googong (before
Australia, ACT 62 5,2,0 c-tm SC dacite & granite good, dry placed
raising)
slate, chert &
Kisenyama Japan 88 5,2,0 c-tm - good (?)
sandstone
Kurokawa Japan 98 5,2,0 c-tn (u) - - good (?)

La Grande 2 (LG-2) Canada, James Bay 160 5,2,0 c-tn (u) SM granite ? reasonable to good, dry placed

Maroon Australia, Qld 52 5,2,2 c-tk SC/CL porphery reasonable to good


Matahina New Zealand 85 5,1,0 c-tm CL/ML ignimbrite poor, dry placed
Mud Mountain USA, Washington 128 5,1,0 c-tm SM/GM andesite & tuff dumped and sluiced
Naramata Japan 158 5,1,0 c-tn GM/GC rhyolite & granite good
Netzahualcoyotl Mexico 132 5,2,0 c-tn MH conglomerate poor to reasonable.

Nillahcootie Australia, Vic. 35 5,2,0 c-tm CL sandstone & mudstone poor (?)

Nottely USA, Tennessee 56 5,1,1 c-tk CL quartzite & mica schist dumped and sluiced

Parangana Australia, Tas. 53 5,2,2 c-tm SM quartzite & schist reasonable to good, dry placed

hornfelsed andesite &


Peter Faust Australia, Qld 51 5,2,0 c-tn CL reasonable to good
andesitic tuff

Round Butte USA, Oregon 134 5,2,0 c-tn (u) SM basalt reasonable to good, dry placed

Rowallan Australia, Tas. 43 5,1,0 c-tn GM quartzite reasonable to good, dry placed

Sagae Japan 112 5,2,0 c-tn - - good (?)


Seto Japan 102 5,1,0 c-tn - - good (?)
Shimokotori Japan 119 5,1,0 c-tn - - good (?)

brecciated quartzites &


Split Yard Creek Australia, Qld. 76 5,2,1 c-tm CL reasonable
greenstone

sandstone, some
South Holson USA, Tennessee 87 5,2,1 c-tk CL dumped and sluiced
shale
Srinagarind Thailand 140 5,2,0 c-tn GC/SC limestone good
Svartevann Norway 129 5,2,0 c-tn (u) GM granitic gneiss reasonable, dry placed
Taisetsu Japan, Hokkaido 86.5 5,1,0 c-tn - - good (?)
Talbingo Australia, NSW 162 5,2,0 c-tm (u) CL rhyolite & porphyry good

Tedorigawa Japan, Honshu Island 153 5,2,0 c-tn GM/GC - good (?)
Chapter 7: Deformation behaviour of embankment dams Page 7.20

Table 7.4: Central core earth and rockfill embankments in the database (continued)
n
Height Embankment Class Rockfill
1
Dam Name Country * H Core Core
n
(m) Class Source Compaction Rating
Size Type
Terauchi Japan, Fukuada 83 5,2,0 c-tm SC/SM lutaceous schist good

Thomson Australia, Vic. 166 5,2,0 c-tm SC sandstone & siltstone good, dry placed

Tokachi Japan, Hokkaido 84.3 5,1,0 c-tm GC keratophyre, 30% slate good (?)

Upper Dam Korea 88.5 5,2,0 c-tn GC tuff & rhyolite poor (?)

Upper Yarra Australia, Vic. 90 5,0,0 c-tk CL sandstone & siltstone poor, dry placed

quartz, granite &


Vatnedalsvatn Norway 121 5,2,0 c-tn SM reasonable, dry placed
gneiss
Waipapa New Zealand 36 5,1,1 c-tk CH/SC ignimbrite poor
Watuaga USA, Tennessee 94 5,2,1 c-tk CL/SC quartzite dumped and sluiced
William Hovell Australia, Vic. 34 5,2,0 c-tm ML rhyodacite reasonable to good
reasonable to good (inner),
Windemere Australia, NSW 67 5,2,1 c-tm CL andesite
reasonable (outer)
Wivenhoe Australia, Qld 59 5,2,1 c-tn CL sandstone good

reasonable to good (inner),


Wyangala Australia, NSW 86 5,2,0 c-tm SC-SM porphyritic gneiss
reasonable (outer), dry placed

n
Note: Class = classification
1
* For Australian dams; Qld = Queensland, NSW = New South Wales, Vic. = Victoria
Tas. = Tasmania, ACT = Australian Capital Territory

Table 7.5: Zoned earth and rockfill embankment case studies


n
Height Embankment Class
1
Dam Name Country * H Comment
Core Core
(m) Classn
Size Type
Andong Korea 83 4,1,1 c-tn SM Rockfill used in outer shoulders
Bradbury USA, California 85 4,1,1 c-tk SC/GC Weathered shale in downstream shoulder
Rockfill in up and downstream shoulder, poorly
Burrendong Australia, NSW 76 4,2,1 c-tk CL
compacted.

Canales Spain 156 4,1,0 c-tn CH Limestone rockfill in up and downstream shoulder

Cardinia Australia, Vic. 86 4,2,1 c-tm (u) SC/CL granodiorite in upstream shoulder

Dixon Canyon USA, Colorado 74 4,1,0 c-vb CL thin outer rockfill zones. Virtually homogeneous.

Eucembene Australia, NSW 116 4,1,0 c-vb SC Quartzite rockfill as thin outer zones.
Greenvale Australia, Vic. 52 4,2,2 c-tk SM granodiorite rockfill in upstream shoulder
Broad GC earthfill zones up and downstream of the
Jackson Gulch USA, Colorado 56 4,0,1 c-tm (u) CL
core, and thin outer rockfill shoulders
poor quality limestone used in up and downstream
La Agnostura Mexico 146 4,0,0 c-tn CL
shoulders
granite and gneiss used in up and downstream
La Grande - LG4 Canada 125 4,1,0 c-tm SM
shoulders
Broad earthfill zones up and downstream of core
Long Lake USA 39 4,0,0 c-tk CL/ML
and thin outer rockfill zones.
Nurek Russia 289 4,2,0 c-tn GC/GM thin outer rockfill zone
Peublo (right
USA, Colorado 51 4,0,1 c-vb CL limited use of rockfill
abutment)
limited rockfill used, outer shoulders. Close to
Rector Creek USA, California 61 4,0,0 c-vb SC/SM
homogeneous classification.
San Justo USA, California 41 4,2,2 c-tk CL limestone rockfill used upstream shoulder
San Luis - Main
USA, California 116 4,2,1 c-vb CL limited rockfill used, outer shoulders
Dam
San Luis - Slide
USA, California 30 to 45 4,1,1 c-vb CL limited rockfill used, outer shoulders
Area

Soyang Korea 123 4,1,0 c-tm GC poor ly compacted rockfill used in outer shoulders

Spring Canyon USA, Colorado 68 4,1,0 c-vb CL limited rockfill used, outer shoulders
dumped and sluiced granite rockfill in upstream
Tooma Australia, NSW 67 4,1,2 c-tk SM
shoulder
limited rockfill used, at up and downstream toe
Tullaroop Australia, Vic. 41 4,1,0 c-tk CL
regions. Could also classify as 2,1,0.
Note: Classn = classification
*1 For Australian dams; NSW = New South Wales, Vic = Victoria
Chapter 7: Deformation behaviour of embankment dams Page 7.21

Table 7.6: Zoned earthfill embankment case studies


Height Embankment Class n
Dam Name Country H Core Core Comment
(m) Class n
Size Type
Benmore New Zealand 110 3,2,1 c-tk GM/GC thick core with compacted gravel shoulders
Cairn Curran Australia, Vic. 44 3,0,0 c-vb CL/ML embankment with very broad core
wet placed core supported by shoulders of well-
Carsington England 36 3,0,1 c-tm CH compacted weathered mudstones. Failed during
construction.
thick core with gravel upstream shoulder and talus
Cobb New Zealand 35 3,0,1 c-tk GM/GP
downstream shoulder
upstream sloping medium width core supported by
Davis USA, Arizona 60 3,0,1 c-tm (u) CL
broad silty sand to silty gravel earthfill zones.
Deer Creek USA, Utah 71 3,0,0 c-tk CL-GC central core of thick width with earthfill shoulders
Granby USA, Colorado 88 3,0,0 c-vb SC-GC embankment with very broad core
Hirakud India 59 3,0,1 c-vb SC-GC embankment with very broad core
Horsetooth USA, Colorado 48 3,0,0 c-vb CL embankment with very broad core

Kastraki Greece 95 3,1,0 c-tn CL thin clay core with well compacted gravel shoulders

embankment with very broad core, upstream gravel


Khancoban Australia, NSW 18 3,0,2 c-vb SM
shoulder.

Kremasta Greece 165 3,1,0 c-tn CL thin clay core with well compacted gravel shoulders

Mammoth Pool USA, California 113 3,2,2 c-tk SM thick core with compacted earthfill shoulders
Medicine Creek USA, Nebraska 48 3,0,0 c-vb CL embankment with very broad core
Meeks Cabin USA, Wyoming 57.5 3,1,1 c-tk CL thick core with gravel shoulders

Navajo USA, New Mexico 123 3,1,1 c-tk CL thick core with compacted gravel and earthfill shoulders

central core of thick width with filter / transition zones up


O'Sullivan USA, Washington 64 3,1,0 c-tk SM
and downstream of the core
Pukaki New Zealand 76 3,2,1 c-tk GM thick core with compacted gravel shoulders

Rosshaupten Germany 41 3,1,1 c-tm GM/ML medium width core with compacted gravel shoulders

Ruataniwha New Zealand 50 3,2,1 c-tk GM thick core with well compacted gravel shoulders
Soldier Canyon USA, Colorado 70 3,0,0 c-vb CL embankment with very broad core
Steinaker USA, Utah 49 3,0,0 c-vb CL embankment with very broad core
Trinity USA, California 164 3,0,0 c-tk SM/GM thick core with gravel shoulders
n
Note: Class = classification
*1 For Australian dams; NSW = New South Wales, Vic = Victoria

Table 7.7: Earthfill embankment case studies


Height Embankment Classification
Name Country H Core Comment
(m) Dam Type Classn
Type
USA, South
Belle Fourche 35 Homogeneous Earthfill 0,0,0 CL-CH concrete facing on upstream slope
Dakota
Biet Netufa Israel 11 Earthfill with rock toe 2,0,0 CH
Earthfill with foundation Foundation filter under downstream
Bonny USA, Colorado 44 1,0,1 ML-SM
filter shoulder
Earthfill with thin rockfill
Enders USA, Nebraska 31.5 2,0,0 SM/ML virtually a homogeneous embankment
shell
Fresno USA, Montana 24 Earthfill with rock toe 2,0,0 ML-CL
USA, North Earthfill with rock/gravel
Heart Butte 42 2,0,0 SC-SM
Dakota toe
Mita Hills Africa, Zambia 49 Earthfill with filters 1,2,1 CL-ML (?) upstream sloping dual chimney filter
Peublo (left thin downstream outer shoulder zone of
1,0,1
abutment USA, Colorado 37 Earthfill, rockfill toe CL gravels. Stability berm on downstream
(or 3,0,1)
embankment) shoulder, placed 1981.
Part earthfill part concrete gravity &
Roxo Dam Portugal 34 Earthfill, rockfill toe 2,0,0 CL/SC
buttress dam.
central chimney filter (dual filters) and
Sasumua Kenya 35 Earthfill with filters 1,2,0 MH large rockfill zones at up and
downstream toe.
Chapter 7: Deformation behaviour of embankment dams Page 7.22

Table 7.8: Puddle core earthfill embankment case studies


Core
Puddle Material
Height Dimensions Select
Year
Name Country H Top Core Zone General Earthfill
Comp.
(m) Width Slope Source Classn Yes/No
(m) (H to V)
hard mudstone - 450 mm layers,
Burnhope UK 1935 41 nk nk Boulder Clay CL Yes
compacted by steam rollers
Challocombe UK 1944 15 1.2 1 to 15 nk ML Yes nk
Cwnwernderi UK 1901 22 1.5 1 to 10 nk CL No ? nk
Dean Head UK 1840 19 nk nk nk nk nk nk
Residual clays (CL) - 150 mm layers
Happy Valley Aust. 1896 25 2.44 1 to 8.25 residual CL Yes compacted by wagons, carts and grooved
rollers
Upper Lias Sand and sand-rock, placed in thin layers
Hollowell UK 1937 13.1 2 1 to 10 nk Yes
blue clay by dozer
Weathered mudstones & sandstones, mix
Holmestyes UK 1840 25 2 1 to 20 Boulder Clay nk No sandy silty clay with gravel to boulder size
fragments
alluvial / fluvial (downstream only), SC/SM,
alluvial /
Hope Valley Aust. 1872 21 1.8 1 to 12 CH Yes placed in 1200 mm layers compacted by
fluvial
3.5t sheepsfoot roller
Ladybower UK 1945 43 nk nk nk nk nk nk
Not sure, possibly weathered shale,
Langsett UK 1904 33 nk nk nk Clay Yes placed in 900 mm lifts and compacted by
slewing of the rails
Weathered mudstones & sandstones, mix
Ogden UK 1858 25 2 1 to 10 Boulder Clay CH No sandy silty clay with gravel to boulder size
fragments
Weathered mudstones & sandstones, mix
Ramsden UK 1883 25 3 1 to 12 Boulder Clay ML/MH Yes sandy silty clay with gravel to boulder size
fragments

Boulder Clay (CL) sandy clay, placed in


Selset UK 1959 39 1.5 1 to 20 Boulder Clay CL Yes
225 mm layers by dozers and 13.6 t rollers
Weathered mudstones & sandstones, mix
Walshaw
UK 1907 22 2.6 1 to 12 Boulder Clay CL No sandy silty clay with gravel to boulder size
Dean
fragments
Widdop UK 1878 20 nk nk nk nk No nk
alluvium, CL/CH clays upstream and
ML/CL sandy silts and silty clays
Yan Yean Aust. 1857 9.6 3.05 1 to 3.3 alluvium CL/CH No
downstream, placed in 600 to 1200 mm
lifts
Weathered mudstones & sandstones, mix
Yateholme UK 1872 17 nk nk Boulder Clay CL No sandy silty clay with gravel to boulder size
fragments
n
Note: Class = classification Year Comp. = year of completion of construction
nk = not known
*1 For Australian dams; NSW = New South Wales, Vic = Victoria

The deformation behaviour has been sub-grouped into deformation during construction
(this section, Section 7.4) and deformation post construction (Sections 7.5 and 7.6). The
deformation behaviour during first filling is discussed as a separate section (Section 7.5)
but the plotted deformation behaviour has, in most cases, been incorporated into plots of
the post construction deformation behaviour presented in Section 7.6.
The analysis of the deformation behaviour during construction of earthfill and earth-
rockfill embankments is mostly concentrated on the central earthfill core zone of the
embankment because this is where instruments are located. Most of the deformation
data collected and analysed is vertical deformations, usually from internal vertical
settlement gauges (IVM) installed in the core as construction proceeds. Data on
Chapter 7: Deformation behaviour of embankment dams Page 7.23

horizontal deformations, usually from internal horizontal movement gauges (IHM) or


inclinometers, is less frequently monitored, but has been collected and analysed for a
number of embankments.
Analysis of deformation requires consideration of the stress conditions imposed
during construction and the stress-strain relationship of the materials used in
embankment construction. In zoned embankment construction it is also important to
consider both the total and effective stress conditions, and the interaction between the
different material zones in the analysis. While finite element analysis can be used for
individual dams, this was impracticable for this study due to the number of case studies
analysed. Therefore, simplified approaches have been used with reference to finite
difference or finite element analysis to indicate the generalised stress conditions and
deformation behaviour.

7.4.1 STRESSES DURING CONSTRUCTION

7.4.1.1 Stresses in a Homogeneous Embankment on a Rigid Foundation


For the simplified case of a homogeneous embankment on a rigid foundation and
assuming elastic behaviour, the vertical and lateral stresses within the embankment are
dependent on the slope geometry, and the total unit weight and Poisson’s ratio of the
earthfill.
During the initial stages of embankment construction the embankment width is
much greater than the height of earthfill. Under these conditions, assuming plane strain
conditions along the dam axis, the stresses below the dam axis can be modelled
assuming zero lateral strain normal to the dam axis (Figure 7.1a). The elastic stress-
strain relationships for a soil element under the embankment axis following placement
of an incremental layer of distributed load p kPa are approximated as:

εx =ε y =0

ν ∆σ z
∆σ z = p ; ∆σ x = ∆ σ y =
(1 − ν )

constrained tangent modulus, ∆D = ∆ σ z / ∆ε z


Chapter 7: Deformation behaviour of embankment dams Page 7.24

∆D (1 + ν ) (1 − 2ν )
Young’s tangent modulus, ∆E =
(1 − ν )

where ε = strain, σ = stress, x, y and z are axis co-ordinates, ν is Poisson’s ratio and ∆
represents the incremental or tangential component of the parameter.
Because of geometry effects, this approximation is not sufficiently accurate for
estimating the change in stresses in a soil element low in the dam in the latter stages of
construction where the layer width is small compared to the constructed embankment
height (Figure 7.1b).

Figure 7.1: Embankment construction indicating (a) broad layer width to embankment
depth ratio in the early stages of construction and (b) narrow layer width to embankment
depth ratio in the latter stages of construction.

To facilitate analysis of the data, total vertical stress profiles during and at the end of
construction were estimated for a variety of symmetrical embankment shapes by elastic
finite difference analysis. Figure 7.2a presents a typical total vertical stress profile at
Chapter 7: Deformation behaviour of embankment dams Page 7.25

end of construction for an embankment with slopes of 1.8H to 1V and Figure 7.2b the
ratio of total vertical stress at the end of construction for various embankment slopes
compared to the total vertical stress calculated simply by the depth below the ground
surface and the unit weight (i.e. γH). Lines of best fit were obtained using a second
order polynomial equation according to Equation 7.3. Values of the coefficients A1 and
A2 for the various embankment slopes analysed are given in Table 7.9.

(
σ z , EOC = γ ⋅ H A1 (h H )2 + A2 (h H )) (7.3)
= Z1 ⋅ γ ⋅ H

where s z,EOC = total vertical stress at end of construction (at depth h/H), ? = bulk density,
H = embankment height, h = depth below crest, A1 and A2 are coefficients and Z1 is the
vertical stress ratio.
For estimation of stresses within the embankment during construction finite element
or finite difference methods can be used. Alternatively, published charts such as those
by Clough and Woodward (1967) or Poulos et al (1972) can be used.

7.4.1.2 Stresses in Zoned Embankments


Zoned embankments present a greater degree of complexity for analysis than the
homogeneous embankment case because of the potential for stress transfer or shedding
of stresses from the core to the shoulder or vice versa. Simplification to a homogeneous
analysis is only acceptable where arching has a negligible influence on the vertical
stress profile. As will be demonstrated in Section 7.4.1.3 this is a reasonable
assumption for:
• Zoned embankments with broad central earthfill zones, e.g. for central core widths
with a combined core slope (upstream and downstream core slope combined)
greater than about 1.5 to 2H to 1V. It may also be a reasonable simplification for
combined central core widths down to 1H to 1V where the well-compacted core has
been placed on the dry side of Standard optimum moisture content, giving a core of
high undrained strength, so that the stress conditions in the core are maintained in
the elastic region.
• Zoned embankments where the compressibility properties of the central earthfill
zone and gravelly or rockfill shoulders are similar; i.e. compacted sandy and
Chapter 7: Deformation behaviour of embankment dams Page 7.26

gravelly soils with non-plastic fines or low (less than about 20% finer than 75
micron) plastic fines contents, with shoulders of compacted gravels or rockfill. This
is on the proviso that pore water pressures generated in the core during construction
are small and potential plastic type core deformations due to lateral spreading of the
core are therefore negligible.

b) 1

Embankment slope angle


Flat
0.8 3H to 1V
Vertical stress ratio, Z 1

2.3H to 1V
1.8H to 1V
0.6 1.5H to 1V
1.3H to 1V

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
Ratio h/H

Figure 7.2: (a) Total vertical stress (σz,EOC ) contours at end of construction for 1.8H to
1V embankment slopes, and (b) vertical stress ratio, Z1, versus h/H under the
embankment axis for various embankment slopes.

Table 7.9: Multiplying coefficients for the calculation of the total vertical stress at the
embankment centreline at end of construction (Equation 7.3)
Embankment Slope Angle A1 A2
1.3H to 1V 0.0084 0.803
1.5H to 1V -0.0098 0.843
1.8H to 1V -0.029 0.886
2.3H to 1V -0.044 0.931
3.0H to 1V -0.062 0.978
Chapter 7: Deformation behaviour of embankment dams Page 7.27

Where the core is of thin to medium width and the compressibility properties of the core
and shoulders are dissimilar it is not possible to use the simple methods described in
Section 7.4.1.1 to analyse the deformation of the core. The problem is no longer one-
dimensional because the vertical strain of the core under the embankment centreline
comprises two components, one due to the vertical strain under the imposed vertical
stress and a second component due to the lateral strain of the core. In addition, stress
transfer effects cannot be ignored. Meaningful analysis of stresses and strains during
construction can only be undertaken using finite element modelling based on
constitutive models that reasonably approximate the as placed properties of the various
materials used in construction.
However, to facilitate the analysis of the case study deformation data, the analysis of
a simple plane strain finite difference model of a zoned embankment was undertaken to
assess the general stress conditions and trend of deformation behaviour during
construction. The modelled embankment (Figure 7.3) consisted of a 100 m high central
core earth and rockfill dam, of thin core width, on a rigid foundation. Construction was
modelled in ten lifts each of 10 m height. A Mohr-Coulomb linear-elastic perfectly
plastic model was used for both the central core and rockfill. Three analyses were
carried out each using the same properties for the rockfill (Figure 7.3) and varying the
properties of the core material (Table 7.10) as follows:
• Case 1 – core with similar shear strength properties to the rockfill. Poisson’s ratio
was assumed to be the same for the core and rockfill shoulders. This case is
considered typical of a well-compacted silty gravel core with non-plastic fines.
• Case 2 – A very stiff core with Poisson’s ratio much greater than that of the rockfill
(0.40 compared to 0.22) and strength indicative of the undrained strength envelope.
• Case 3 – A wet placed core. The core is assumed to behave nearly fully undrained
with a Poisson’s ratio of 0.48 and a small equivalent undrained friction angle of 5
degrees.

The same modulus has been used for the core (100 MPa) and the rockfill (80 MPa)
in all three cases to minimise where possible the number of variants. A friction angle of
45 degrees is used to model the rockfill.
The results, shown in Figure 7.4 and Figure 7.5, highlight the influence of the
difference in shear strength and Poisson’s ratio of the core on the stresses within the
embankment and on the deformation behaviour. Where the properties of the core are
Chapter 7: Deformation behaviour of embankment dams Page 7.28

similar to those of the rockfill (Case 1) the lateral stress distribution is similar to the
homogeneous case and lateral displacements in the core and rockfill are small (Figure
7.4 and Figure 7.5c). The vertical stress distribution on the embankment centreline for
this case (Case 1) is similar to the homogeneous case, except that the higher modulus
core attracts higher stresses during construction. The assumption of homogeneous
conditions reasonably models the lateral stresses but will under-estimate the vertical
stresses in the core where the core is less compressible than the rockfill, and vice versa
if the core is more compressible than the rockfill. For broader core widths the vertical
stress profile on the centreline approaches the homogeneous case (refer Section 7.4.1.3).
Where the stress-strain properties of the core are different to those of the rockfill,
such as for Cases 2 and 3, the stress profiles (Figure 7.4) differ significantly from the
homogeneous case. Higher lateral stresses are developed during construction as well as
arching across the core, as shown. Outward lateral displacements of the core and
rockfill are observed that are much larger than for Case 1 or for the homogeneous case
(Figure 7.4 and Figure 7.5c). The displacements are caused by the differential lateral
stress conditions that develop between the core and shoulders. For Case 3, yield
conditions are reached in a large portion of the core and in this case the rockfill is in
effect acting as support for the core.
The results also show:
• Lateral stresses and strains increase with increasing Poisson’s ratio of the core,
reflecting yielding and plastic deformation of the core, particularly for Case 3.
• Maximum lateral strains are observed at a depth below the crest of approximately
65% of the embankment height (h/H = 0.65).
• The lateral strains, which occur largely as plastic deformation of the core, contribute
significantly to the vertical strain observed in the core.
• Arching occurs in the narrow core as a result of the large differential vertical
deformation between the core and the shoulders. Figure 7.5a shows the vertical
stress profile at end of construction for Case 3 is much less than for the
homogeneous condition, and Cases 1 and 2.

The finite difference analysis was based on the assumption of plane strain
conditions. This assumption is valid for embankments that are long compared to their
height because strains in the direction parallel to the dam axis (i.e. y-direction in Figure
7.1) are negligible compared to those measured vertically and perpendicular to the dam
Chapter 7: Deformation behaviour of embankment dams Page 7.29

axis. Consideration should be given to the potential for cross-valley arching and its
effect on the stresses within the embankment for embankments constructed in steep
valleys with narrow river sections.

Figure 7.3: Embankment model for finite difference analysis during construction.

Table 7.10: Properties of central core used in finite difference analyses


Modulus, Poisson’s Bulk Unit Cohesion, Angle of
E ratio, ν Weight, γ c Internal
Analysed Case
Friction, φ
(MPa) 3
(kN/m ) (kPa) (degrees)
Case 1 – zoned embankment with 100 0.22 20 0 45
core similar to rockfill
Case 2 – zoned embankment with 100 0.4 20 75 20
very stiff core
Case 3 - zoned embankment with 100 0.48 20 30 5
stiff, wet core
Homogeneous (same as rockfill) 80 0.22 20 0 45
Cases 4 to 7 – zoned embankment, 40 0.33 20 2 32
core slope 0.2H to 1V to 0.75H to
1V (refer Section 7.4.1.3)
Chapter 7: Deformation behaviour of embankment dams Page 7.30

Figure 7.4: Total vertical stress, total lateral stress, and lateral (horizontal) displacement distributions at end of construction for zoned earthfill
finite difference analysis.
Chapter 7: Deformation behaviour of embankment dams Page 7.31

a) 1

σv = γ.H
Case 1 - zoned, core similar to rockfill
0.8
Case 2 - zoned, very stiff core
Ratio of vertical stress, Z1 .
(at end of construction) .
Case 3 - zoned, stiff wet core
Homogeneous
0.6

0.4

0.2 Note: Total vertical stresses plotted as a


ratio of γ.H.
h = depth below crest

0
0 0.2 0.4 0.6 0.8 1
Depth ratio h/H

b) 0

10

20

30
h = depth below crest (m)

40

50

60

70

80
Case 1 - zoned, core similar to rockfill
Case 2 - zoned, very stiff core
90
Case 3 - zoned, stiff wet core
Homogeneous
100
0 200 400 600 800 1000 1200
Settlement (at depth h) on the dam axis at end
of construction (mm)
Chapter 7: Deformation behaviour of embankment dams Page 7.32

c) 350
Case 1 - zoned, core similar to rockfill
Case 2 - zoned, very stiff core
Case 3 - zoned, stiff wet core

Lateral deformation at Point a during construction (mm) .


300 Homogeneous

250

200

150

100

50

0
30 40 50 60 70 80 90 100
Constructed height of dam (m)

Figure 7.5: Finite difference modelling results; (a) total vertical stress on the dam
centreline at end of construction, (b) settlement profile at dam axis at end of
construction, and (c) lateral deformation at Point a (Figure 7.3) during construction.

7.4.1.3 Arching or Stress Transfer in Zoned Embankments

Arching across the core of zoned embankments, including puddle core earthfill
embankments, has important implications on the deformation behaviour of the core
during construction as well as the potential for hydraulic fracture post construction due
to the reduced stresses in the core. Bishop (1952), for undrained conditions in puddled
core dams, and Nonveiller and Anagnosti (1961), for fully drained conditions in zoned
embankments, have developed methods for assessment of arching for embankments
with narrow cores. The progression to finite element analysis and improvements in
constitutive models to more accurately model the stress-strain relationship of materials
allows for realistic assessment of arching in zoned embankments, as shown by the
analysis of Beliche dam (Naylor et al 1997) (Figure 7.6a) and the modelling of puddle
core earthfill dams by Dounias et al (1996) (Figure 7.6b).
Chapter 7: Deformation behaviour of embankment dams Page 7.33

Figure 7.6: Finite element analysis of zoned embankments highlighting arching of core
zone from (a) Beliche dam at about mid-height, elevation 22.5 m (Naylor et al 1997),
and (b) a puddle core embankment (Dounias et al 1996).

As the published analyses and those from Section 7.4.1.2 show, arching is developed in
zoned embankments with narrow cores not only due to the core being more
compressible than the supporting shoulders but also due to lateral spreading of the core.
Bishop and Vaughan (1962) comment that for the Selset puddle core earthfill
embankment the vertical deformations of the core were largely a result of plastic
deformations due to lateral spreading and not consolidation type settlements due to pore
water pressure dissipation.
Factors contributing to the potential for arching are the difference in strength and
compressibility properties between the core, filter/transition and outer earth or rockfill
zones, the stress path and the embankment geometry (particularly the width of the core).
Some examples of total vertical pressures measured during construction in the core
of zoned embankments using pressure cells are shown in Figure 7.7, and are compared
to the total vertical stress estimated for a homogeneous embankment allowing for the
embankment shape. The core width to depth ratio is 0.4 to 0.5 for most of the cases (i.e.
Chapter 7: Deformation behaviour of embankment dams Page 7.34

thin cores) and is 0.7 for Talbingo dam. The near symmetrical thin cores comprised
sandy clays to clayey sands and clayey gravels placed close to or on the wet side of
Standard optimum moisture content and were supported by reasonably to well
compacted filters, and gravel or rockfill shoulders.
For Kastraki, Thomson, Beliche and Chicoasen dams the measured vertical stresses
are less than the homogeneous case. All four cases are considered to represent arching
across the core. At the higher stress levels associated with the latter stages of
construction the measured stresses tend toward the horizontal (i.e. minimal measured
increase in total stress with increasing embankment height) at Kastraki, Thomson and
Beliche dams. At Chicoasen dam (Alberro and Moreno 1982) the pressure cell
measurements do not show the curvature in the latter stages of construction as do the
other three cases.
For Talbingo dam the pressure cells in the mid region of the slightly upstream
oriented core show a linear stress response with the stresses plotting close to the
homogeneous case indicating arching is not a factor. However, the pore water pressure
response in the core and settlement during construction (refer Appendix G, Section
1.14) indicate that arching occurred in the vicinity of the near vertical downstream core
/ filter interface but not in the central region or the upstream region next to the core /
filter interface sloped at 0.9H to 1V.

3000
Talbingo Dam (h/H = 0.97)
Measured total vertical stress (kPa) .

Talbingo Dam (h/H = 0.81)


2500 Chicoasen Dam (h/H = 0.81)
Beliche Dam (h/H = 0.76)
Thomson Dam (h/H = 0.95)
2000 Thomson Dam (h/H = 0.71)
Kastraki Dam (h/H = 0.70)
Homogeneous Dam
1500

1000

500 Note: h/H refers to the location of the pressure cell in the core,
where H = embankment height, h = depth below crest

0
0 500 1000 1500 2000 2500 3000 3500 4000
Total vertical stress estimated assuming a homogeneous dam condition (kPa)
Figure 7.7: Estimated versus measured (from pressure cells) total vertical stresses
during construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.35

Further numerical analysis was undertaken to analyse the effect of core width (in zoned
embankments) on the stresses during construction. The same procedure and
embankment geometry was used as described in Section 7.4.1.2 analysed with
symmetrical core slopes of 0.2H, 0.3H, 0.5H and 0.75H to 1V (Cases 4 to 7
respectively). Properties for the core are given in Table 7.10 and those for the rockfill
were as per the earlier analysis (Figure 7.3). For these cases the core was modelled with
a Young’s modulus of 40 MPa, half that of the rockfill. The results are presented in
Figure 7.8.
The total vertical stresses on the embankment centreline at Point b (Figure 7.3, h/H
= 0.70) for each stage of construction are presented in Figure 7.8c, and include those
from Cases 2 and 3. The total vertical stress for the zoned cases initially approximates
that for the homogeneous case at low stress levels and then curves away at higher stress
levels, similar to that observed for the thin core case studies (Figure 7.7) and is
indicative of arching of the core.
The results of the analysis also show that at end of construction the total vertical
stresses on the embankment axis increase with increasing core width, and for this
example, approached stresses close to the homogeneous case for core slopes greater
than about 0.5H to 1V (i.e. thick cores). However, there is still a significant differential
in stress at the interface between the core and rockfill as can be seen in Figure 7.8b.
Chapter 7: Deformation behaviour of embankment dams Page 7.36

c) 1400
Case 4 - 0.2H to 1V
Total vertical stress at Point b (kPa) . 1200
Case 5 - 0.3H to 1V
Case 6 - 0.5H to 1V
from numerical analysis .
Case 7 - 0.75H to 1V
1000 Case 2 - very stiff core
Case 3 - stiff, wet core
800

600

400

200
Note: H to V refers to the core slope (refer Figure 7.3)
Point b located on axis at h/H = 0.70 (Figure 7.3)
0
0 200 400 600 800 1000 1200 1400
Total vertical stress for homogeneous case (kPa)

Figure 7.8: Total vertical stress profiles from the numerical analysis of construction for
zoned embankments with different core slopes; (a) 0.2H to 1V case, (b) 0.5H to 1V
case, and (c) total vertical stress at h/H = 0.70 with increasing embankment height.

7.4.1.4 Pore Water Pressures Developed During Construction


So far the discussion on stresses during construction has been on total stresses.
However, for a more complete assessment of the deformation behaviour it is important
to consider effective stresses and the effective stress paths during construction. For
permeable embankment materials such as clean or free draining rockfills, gravels and
sands the assumption of drained conditions during construction is valid. Therefore
effective stresses and the effective stress state in these materials are readily assessable
from the total stress conditions and consideration of partial impoundment, if this occurs
during construction.
For low permeability embankment materials effective stresses are more difficult to
predict prior to construction, but their assessment can be critical in terms of
embankment stability. Therefore, instrumentation is often installed and monitored
during construction in the form of piezometers, pressure cells, and internal and external
deformation monitoring gauges. However, due to the unreliability of pressure cells,
total stresses are generally assessed from finite element analysis, which can also be
unreliable.
At placement, earthfill materials are partially saturated, with the degree of saturation
dependent on the moisture content, material type, and the method and degree of
compaction of the earthfill. The pore water pressures that are developed in the earthfill
Chapter 7: Deformation behaviour of embankment dams Page 7.37

during construction, assuming it is well-compacted, are dependent on a number of


factors including the initial degree of saturation, material compressibility properties,
permeability, time of construction and applied stress levels. It is convenient to consider
the positive pore water pressures in terms of ru, the pore water pressure coefficient, as
defined in Equation 7.4.

ru = u σ z or ∆ru = ∆u ∆σ z (7.4)

where ru = pore water pressure coefficient, u = pore water pressure, σz = total vertical
stress and ∆ represents the incremental or tangential component of the parameter. The
term ru is used for the pore water pressure coefficient determined from case study data
rather than B , as defined in Equation 7.5 (on page 7.40), because the lateral stresses (σx
and σy) parallel and perpendicular to the dam axis may not be equal.
Three main types of pore water pressure response are generally observed (Figure
7.9):
• Negligible or limited positive pore water pressure response (ru generally less than
0.1 to 0.2). Gould (1953, 1954) observed and the case study data indicates that for
earthfills placed drier than about 0.5% to 1% of Standard optimum limited positive
pore water pressures are developed during construction for most soil types.
• Moderate to very high sustained positive pore water pressure response ( ∆ ru greater
than 0.5) with negligible to very limited drainage. This type of response is generally
observed for clayey sands, clayey gravels and dominantly clayey earthfill types
placed wetter than about 0.2 to 0.5% dry of Standard optimum (i.e. close to or above
Standard optimum). Often the positive pore water pressure response is preceded by
a period of negative pore water pressure response at low stress levels (Figure 7.9).
• Drainage during construction is significant. This is generally observed for earthfills
of silty sands to silty gravels placed close to or wet of Standard optimum (i.e. wetter
than about 0.2 to 0.5% dry of Standard optimum), but in some cases it is also
observed for clayey sands and clayey gravels with low plasticity and possibly low
content fines (less than about 15 to 25% finer than 75 micron). The typical pore
water pressure response for these earthfills when placed wet of optimum is for an
initially high pore water pressure coefficient, which then decreases with increasing
embankment height above the piezometer. Partial drainage during shutdown is also
evident.
Chapter 7: Deformation behaviour of embankment dams Page 7.38

There are case studies that fall outside or between these general types of pore water
pressure response.

1600
sustained high pore water
1400 pressure response
Pore Pressure (kPa)

1200
negative pore water
1000 pressure response at low partially drained pore
stresses not shown water pressure response
800

600

400
negligible pore water
200 pressure response

0
0 500 1000 1500 2000 2500 3000
Vertical Stress (kPa)
Figure 7.9: Idealised types of pore water pressure response in the core during
construction.

Hilf (1948) provides a method for prediction of the pore water pressure response for
partially saturated soils derived from one-dimensional consolidation theory and applied
to field prediction using oedometer tests on laboratory samples prepared at proposed
field placement specifications. The method assumes undrained conditions, and field
compression is one-dimensional (i.e. zero lateral strain) and due to compression of the
air voids. The method gives predictions indicating that the rate of pore water pressure
increase becomes very rapid as the air voids volume approaches zero and a saturated
condition is reached. Once saturated the pore water pressure increase is assumed to be
equal to the increase in total vertical stress (i.e. no change in effective stress). Fredlund
and Rahardjo (1993) comment that whilst the Hilf method provides a reasonable
estimate of pore water pressure, it will over-estimate it because the derivation assumes
matric suction equals zero (i.e. the pore water pressure equals the pore air pressure).
Khalili (2002) comments that for partially saturated soils of very low permeability
(i.e. where undrained conditions may be a reasonable assumption) the rate of pore water
pressure increase in confined compression will always be less than the rate of vertical
stress increase due to the presence of a small volume of air voids that remain under the
magnitude of effective stresses likely in embankment dams. The field data for clayey
earthfills of very low permeability placed close to or above Standard optimum moisture
content tend to support this (Figure 7.10), with ma ximum rates of pore water pressure
Chapter 7: Deformation behaviour of embankment dams Page 7.39

increase equivalent to 60 to 90% of the change in total vertical stress (i.e. ∆ ru = 0.6 to
0.9).

2500
Talbingo dam - Piezo HP8
Dartmouth dam - Piezo 31
2000 Blowering dam - Piezo 16
Pore Pressure (kPa) .

Thomson dam - Piezo 7


Kastraki Dam
1500

∆ u = ∆σ z
1000

500

0
0 500 1000 1500 2000 2500 3000
Measured / Estimated Vertical Stress (kPa)

Figure 7.10: Pore water pressure response during construction in the core zone of clayey
earthfills placed at close to or wet of Standard optimum moisture content.

It is noticeable from Figure 7.10 that the pore water pressure coefficient, ∆ ru, may
decrease toward the end of construction. The reason for this is possibly partly due to
pore water pressure dissipation. A more likely explanation is that for piezometers in the
mid to lower region of the core the total stress ratio ( ∆σ 3 / ∆σ 1 ) in Skempton’s
equation for pore water pressure (Equation 7.5) is not constant but decreases in the latter
stages of construction where the layer width is small compared to the height above the
gauge (Figure 7.1b). Assuming the coefficients A and B are constant, the effect of a
reduction in total stress ratio would be for a reduction in the ratio ∆u / ∆σ 1 .
In contrast, the pore water pressure observations in the narrow, medium to high
plasticity wet placed clay core at Yonki dam indicated that the pore water pressure
actually increased during non-construction periods (Fell et al 1992). This behaviour is
possibly due to yielding as a result of the strain rate dependency of the limit state
surface. Whilst not common, an increasing pore water pressure response during non-
construction periods has been observed in the soft foundations of fills constructed on
soft ground, several examples of which are referred to in Section 4.4.2 of Chapter 4.
For wet placed earthfills of permeability and construction staging such that partial
drainage during construction occurs, the pore water pressure response displays an
interesting feature that is replicated in laboratory testing. Triaxial isotropic compression
Chapter 7: Deformation behaviour of embankment dams Page 7.40

tests on wet placed silty sands and silty gravels (Bishop 1957; Bernell 1982) undertaken
in stages of undrained loading with a period of drainage between loading stages (Figure
7.11) show a reduction in Skempton’s pore water pressure parameter, B (Equation 7.5),
during subsequent stages of undrained loading.
A similar pore water pressure response is observed in wet placed dominantly sandy
and gravelly soils (silty sands and gravels, and some clayey sands and gravels) where
partial dissipation of pore water pressures occurred in shutdown periods during
construction (Bishop 1957; Eisenstein and Law 1977; Nakagawa et al 1985). Bishop
(1957) considers the response is due to a decrease in the compressibility of the earthfill
under increased effective stress conditions that result from partial drainage.
A decrease in the degree of saturation during dissipation of pore water pressures at
shutdown periods could also influence the pore water pressure response afterward. The
relatively high degree of saturation at placement is likely to result in the migration of
water rather than air from the soil structure during the drainage stage, and as pressures
dissipate previously dissolved air in water is likely to return to the air phase. The net
effect could result in an overall decrease in the degree of saturation below that at initial
placement.

∆u   ∆σ 3 
= B = B 1 − (1 − A )  1 −  (7.5)
∆σ 1   ∆σ 1 

where u = pore water pressure, σ1 = major principal stress, σ3 = minor principal stress,
∆ represents incremental or tangential component of the parameter, and B , B and A are
pore water pressure parameters defined by Skempton (1954).
Chapter 7: Deformation behaviour of embankment dams Page 7.41

Figure 7.11: Triaxial isotropic compression tests with staged undrained loading and
partial drainage on a partially saturated silty sand (Bishop 1957).

7.4.1.5 Summary of Stress Conditions During Construction


The total and effective stress conditions established in an embankment during
construction are shown to be dependent on the embankment geometry, the embankment
zoning geometry, and the strength and compressibility properties of the embankment
materials.
For the simplest case of a homogeneous embankment on a rigid foundation, elastic
solutions show that under the embankment centreline the vertical stress profile is
dependent on the embankment geometry and the lateral stresses are dependent on the
Poisson’s ratio of the earthfill and embankment geometry. During the initial stages of
construction when the embankment width is large in comparison to its height the
assumption that the increase in stress equals the depth of soil times its weight is
reasonable in the central region of the embankment. However, in the latter stages of
construction this assumption is no longer valid due to the decreasing layer width in
relation to the embankment height.
The homogeneous model is also a reasonable assumption for the stress conditions
under the embankment axis for:
• Zoned embankments with broad central earthfill zones where the earthfill has been
well compacted and placed drier than about 0.5% of Standard optimum moisture
content. Under these placement conditions the core has a high undrained strength
and the stress conditions in the core are likely to be maintained in the elastic range.
Analysis and the case study data suggests central core widths with a combined core
Chapter 7: Deformation behaviour of embankment dams Page 7.42

slope (i.e. upstream and downstream core slope combined) greater than about 1.5 to
2H to 1V are sufficiently broad, but it may also be a reasonable simplification for
combined central core widths down to 1H to 1V.
and,
• Zoned embankments where the compressibility properties of the central earthfill
zone and gravelly or rockfill shoulders are similar; i.e. compacted sandy and
gravelly soils with non-plastic fines or low (less than about 20% finer than 75
micron) plastic fines contents, with shoulders of compacted gravels or rockfill. This
is on the proviso that pore water pressures generated in the core during construction
are small and potential plastic type core deformations due to lateral spreading of the
core are therefore negligible.

For zoned embankments the stress conditions during construction become more
complex and are affected by the differential strength and compressibility properties of
the embankment materials and the embankment zoning geometry. Total vertical
stresses are affected by stress transfer or arching not only from differential
compressibility but also from differential lateral stresses and the resulting lateral
deformations to reach equilibrium stress conditions. Lateral stresses are affected by
differences in Young’s modulus and Poisson’s ratio between the core and shoulders.
Finite element analysis is required for analysis of stresses and deformations in zoned
earthfill embankments due to the complexity of the problem. Whilst finite element
analysis is a standard tool for assessment of stresses and deformations during
construction of embankment dams there is often a trade off between accuracy and
simplicity (Duncan 1996) due to the complexity of the constitutive model to accurately
model the strength and compressibility properties of partially saturated earthfills and the
free draining gravel and rockfill zones.
For assessment of the effective stress conditions it is necessary to consider the pore
water pressure response in the partially saturated embankment materials that are not free
draining. Advances in the theory of partially saturated soils have developed constitutive
models and methods that allow better estimation of pore water pressures than the
simplistic Hilf (1948) method, but they have not been studied in detail here. Partial
drainage during construction should also be considered.
Chapter 7: Deformation behaviour of embankment dams Page 7.43

7.4.2 LATERAL DEFORMATION OF THE CORE DURING CONSTRUCTION

Table 7.11 summarises the lateral deformation of the core for 12 central core earth and
rockfill embankments. In all cases the central, well-compacted core is of thin to
medium width and is supported by filters/transition zones and shoulders of rockfill
and/or gravels. Further information on the material and geometrical properties of the
core, filter / transition zones and outer shoulders is given in Table 7.12 and the tables in
Section 1 of Appendix F. No data is given for Beliche dam, but Naylor et al (1997)
comment that the deformations during construction distorted the inclinometer tubes
such that lateral deformation measurements could not be taken, indicating the lateral
deformation of the core was likely to have been quite large.
Lateral deformations were, in most cases, measured by internal horizontal
movement gauge (IHM) installed in the filter and rockfill zones downstream of the core,
and in some cases upstream of the core. The displacements quoted in Table 7.11 are
from the IHM gauge closest to the core, usually from within the downstream filter or
transition zone. The “lateral displacement ratio” is defined as the lateral displacement
divided by the core width. Where only deformations on the downstream side of the core
were available the “lateral displacement ratio” of the core has been estimated by
dividing the displacement by the half width of the core and it is assumed that this
displacement ratio is representative of the core at this location.
Figure 7.12 presents the estimated lateral displacement ratio versus fill height above
the measuring gauge at the end of embankment construction. Figure 7.13 presents the
estimated lateral displacement ratio versus fill height above the measuring gauge during
embankment construction. It would be preferable to present these plots of the lateral
displacement ratio versus vertical stress, but for several of the cases no pressure cell
records were available and the likelihood of arching meant that vertical stresses could
not be accurately estimated.
The data shows that the lateral displacement ratios during construction are
significant for zoned embankments with thin to medium width cores. Estimated lateral
displacement ratios range from less than 0.5% up to more than 3% from the case study
data.
Chapter 7: Deformation behaviour of embankment dams Page 7.44

Table 7.11: Lateral deformations of the central core at end of construction for central
core earth and rockfill embankments.
Core Summary
Displacement Details
Dam Height (refer Table 7.12 for more details)
Name (m) Material Type / Width Depth h/H Displacement LDR *3
Placement (H to V) (m) *1 (mm) *2 (%)
Copeton 113 SC – medium plasticity fines, 0.8H to 1V 40 0.35 140 -0.77
spec. = 1% dry to 1% wet of 59 0.52 150 -0.58
OMC (likely dry of OMC) 89 0.79 285 -0.76
Wyangala 86 SC/SM – 1% dry of OMC 0.8H to 1V 51 0.59 67 -0.24
66 0.77 158 -0.70
Fukada 56 GC – no details on moisture 0.6H to 1V 18.5 0.33 -6 0
spec., suspect dry of OMC. 36.5 0.65 42 -0.31
Dartmouth 180 SC/SM – medium plasticity 0.65H to 1V 64 0.36 385 (dn), -450 -1.71
fines, spec. 0.5% dry to 2% (average) 124 0.69 (up)
wet. 1220 (dn) -2.69
Beliche 55 GC – low plasticity fines, 0.44H to 1V - - - large
spec. ± OMC
La Grande, 125 SM - non-plastic fines, spec. 0.5H to 1V 70 0.56 300 (dn), -200 -1.3
LG4 1% dry to 2% wet (up)
Nurek 289 GC/GM – no details on 0.46H to 1V 145 0.50 425 -0.7 to -1.2
moisture spec., suspect at or
above OMC given high PWP.
Thomson 166 SC – medium plasticity fines, 0.5H to 1V 88 0.53 770 (dn), -750 -3.3
2% wet of OMC 118 0.71 (up)
445 -1.45
Gepatsch 153 GM/GC – 0.5 to 2% wet of 0.25H to 1V 82 0.54 300 to 400 (dn - -1.25 to -1.7
OMC 122 0.80 up)
750 (dn - up) -2.2
Talbingo 162 CL – medium plasticity, at 0.7H to 1V 95 0.59 410 (dn), -1310 -2.80
OMC (up)
Blowering 112 CL/SC – medium plasticity 0.9H to 1V 73 0.65 735 -2.26
fines, 0.3% dry to 0.3% wet
El 148 CL/CH - medium to high 0.18H to 1V 74 0.59 - -1.60
Infiernillo plasticity, 3.7% wet of OMC
*1 Depth = depth below crest.
*2 Lateral displacements mainly from IHM gauges and from the gauge closest to the core (usually within
the downstream filter/transition zone). Downstream (dn) is the default. Denoted (up) if from
upstream of the core.
*3 LDR = lateral displacement ratio and is calculated from the measured lateral core displacement
divided by the core width. Values are negative in accordance with the convention that compressive
strains are positive.
Chapter 7: Deformation behaviour of embankment dams Page 7.45

-3.5%
Thomson
Lateral displacement ratio of the core (%) . Clayey cores (CL to CH)

-3.0%
Clayey sand cores (SC)
Non plastic cores (SM or GM) Talbingo
Dominantly gravelly cores (GC or GM) Dartmouth
-2.5%
Blowering
Gepatsch
-2.0%
Dartmouth
El Infiernillo
-1.5% Gepatsch Thomson
La Grande, LG4 Nurek
-1.0%
Copeton Wyangala
Copeton
Copeton
-0.5%
Fukada
Fukada Wyangala
0.0%
0 20 40 60 80 100 120 140 160
Fill height above gauge (m)

Figure 7.12: Estimated lateral displacement ratio of the core at end of construction.

-3.5%
Lateral displacement ratio of the core (%) .

-3.0%

-2.5%

-2.0%

-1.5%

-1.0%

-0.5%

0.0%

0.5%
0 20 40 60 80 100 120 140
Fill height above gauge (m)

Dartmouth (h/H = 0.69) Dartmouth (h/H = 0.36) Wyangala (h/H = 0.59) Wyangala (h/H = 0.77)
El Infiernillo (h/H = 0.59) Thomson (h/H = 0.71) Thomson (h/H = 0.53) Talbingo (h/H = 0.59)
Fukada (h/H = 0.65) Copeton (h/H = 0.79) Blowering (h/H = 0.65)

Figure 7.13: Estimated lateral displacement ratio of the core versus fill height above
gauge during construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.46

Table 7.12: Summary of embankment and earthfill properties for cases used in the analysis of lateral core deformation during construction.
Core Details Filters / Transition Shoulders
Height Zoning 2
Dam Name n 1 Width *
(m) Class * ASCS Fines *4 Plasticity FMC to Pore * 6 Material Width * 7 Compaction Material Compaction Comment
3
(H to V) * (%) *4 OMC * 5 Pressure Rating * 8 Rating * 9
Fukada 56 5,1,2 c-tm 0.6 to 1 GC (?) - - No details - - thin good (?) weathered good (?) relatively high moduli for
sandstone & weathered rockfill
granite “dirty”
La Grande, 125 4,1,0 c-tm 0.5 to 1 SM 30 non-plastic 1% dry to 2% negl to low sandy mod to thin good gravels good (inner) outer rockfill shoulders
LG4 wet, suspect gravels (granite & gneiss),
avg. is wet. reasonable to good
Nurek 289 4,2,0 c-tn 0.46 to 1 GC/GM - - No details med to high sands & mod to thin good gravels good Suspect core placed close
gravels to OMC given high PWP
response.
Copeton 113 5,2,0 c-tm 0.8 to 1 SC 25 to 60 medium Spec. 1% dry negl to low crushed thin good granite (very high reasonable to Likely dry core placement
to 1% wet rock strength) good given low PWP.
Wyangala 86 5,2,0 c-tm 0.8 to 1 SC-SM - - 1% dry negligible sands & mod to thin good porphyritic gneiss reasonable to outer rockfill of
crushed (high to very high good (inner) reasonable to poor
rock strength) compaction.
Thomson 166 5,2,0 c-tm 0.5 to 1 SC 44 medium 2% wet med to high crushed thin reasonable sandstone and good (inner) High to very high strength
sandstone siltstone rock. Outer rockfill of
lesser quality.
Gepatsch 153 5,2,0 c-tn 0.25 to 1 GM/GC 30 - 0.5 to 2% wet low to high gravels broad reasonable to gneiss reasonable clean to “dirty” coarse
(rapid good rockfill.
dissipation)
Beliche 55 5,2,2 c-tn 0.44 to 1 GC 20 to 36 low Spec. ± OMC med to high sandy thin reasonable (?) rockfill (med to poor to inner rockfill high fines
gravels high strength) reasonable and low moduli
Dartmouth 180 5,2,0 c-tm 0.6-0.7 to 1 SC/SM 24 medium Spec. 0.5% high to very crushed thin good granitic gneiss good (inner) reasonable compaction of
dry to 2% wet high rock outer rockfill
Talbingo 162 5,2,0 c-tm (u) 0.7 to 1 CL 18 to 70 medium OMC very high crushed broad good rhyolite & good (inner) reasonable compaction of
rock porphry outer rockfill
Blowering 112 5,2,0 c-tm 0.9 to 1 CL/SC 30 to 80 medium 0.3% dry to med to high sandy thin good quartzite & good (inner), relatively low moduli of
0.3% wet gravels phyllite reasonable rockfill
(outer)
El Infiernillo 148 5,2,0 c-tn 0.18 to 1 CL/CH 54 to 70 med to high 3.7% wet suspect very sands & moderate good diorite & reasonable very narrow core.
high gravels silicified (inner) to poor
conglomerate (outer)
Note: The notes to Table 7.12 are on the next page
Chapter 7: Deformation behaviour of embankment dams Page 7.47

Notes to Table 7.12:


*1 Zoning classification system defined in Figure 1.5 in Section 1.3.3 of Chapter 1.
*2 core width refers to the width of the core compared to the height (H = horizontal, V = vertical)
*3 ASCS = Australian Soil Classification System.
*4 Fines = % finer than 75 micron, Plasticity: low = low plasticity, med = medium plasticity, high = high
plasticity
*5 FMC to OMC = field or placement moisture content relative to Standard optimum moisture content
*6 Pore water pressures during construction; negl = negligible (ru < 0.1), low = ru of 0.1 to 0.25, med =
medium (ru = 0.25 to 0.5), high = ru of 0.5 to 0.75, very high = ru of > 0.75.
7
* Qualitative assessment of width of filters / transition zone; thin = 3 to 6 m wi dth, broad = more than
about 20 m but dependent on dam height
8
* Compaction rating for filters / transition based on placement methods, layer thickness and relative
density.
9
* Compaction rating for rockfill summarised in Section 1.3.3 of Chapter 1.

The earlier finite difference analyses of zoned embankments on a rigid foundation


(Section 7.4.1.2, Figure 7.4 and Figure 7.5) showed that the lateral deformation of the
core was dependent on the development of differential lateral stress conditions between
the core and supporting shoulder zones. The greater the stress difference the greater the
lateral deformation of the shoulders to equilibrate the greater lateral stresses in the core.
The rigid foundation also influenced the lateral stress and displacement profiles, the
maximum lateral displacement being observed at approximately one third of the dam
height above foundation level.
In the analyses the compressibility properties of the supporting rockfill zone were
the same for each case. But, the compressibility properties of the supporting rockfill
zone will also influence the amount of lateral deformation. Under similar differential
stress conditions greater deformations will occur for more compressible supporting
shoulders. The case study data tends to support this as discussed below.
Several trends are evident from the lateral displacement case study data:
• Placement moisture content of the core relative to Standard optimum moisture
content is significant. Low lateral displacement ratios (< 1% at fill heights of 35 to
90 m) are observed for earthfills placed on the dry side of Standard optimum
(Wyangala, Copeton and probably Fukada dams). Higher lateral displacement ratios
(1.0 to 2.2% at 60 m fill height) are generally observed for earthfills placed at or on
the wet side of Standard optimum.
• Low lateral displacement ratios are observed for embankments with broad well-
compacted gravelly zones of likely very high moduli adjacent to the central core
Chapter 7: Deformation behaviour of embankment dams Page 7.48

(Nurek, La Grande LG4 and Gepatsch dams). For these cases the cores are silty to
clayey sands and gravels placed at or wet of Standard optimum.
• Large lateral displacement ratios are observed for clayey earthfills placed at or wet
of Standard optimum supported by shoulders of relatively highly compressible
rockfills (Beliche and Blowering dams). At Beliche dam, Naylor et al (1997)
describe the inner rockfill zone of weathered, medium strength schists and
greywackes as “relatively lightly compacted” and of “relatively high
compressibility”. At Blowering dam the rockfill, although wetted and reasonably
(outer Zone 3B) to well compacted (inner Zone 3A), the vertical secant moduli of
the rockfill was calculated from hydrostatic gauges to be relatively low (20 to 40
MPa for Zone 3A placed in 0.9 m lifts and 10 to 20 MPa for the outer Zone 3B).
Further details on the material properties at Blowering dam are given in Section 6 of
Appendix G.
• Arching effects in the very thin wet placed clay core at El Infiernillo dam are
possibly significant in the relatively low lateral strain measured.
• At Talbingo dam the lateral deformation was significantly greater on the upstream
side of the core (1310 mm) compared to the downstream side (410 mm) at elevation
457 m (h/H = 0.59). Possible reasons for this are the slight upstream orientation of
the core and the much broader width of the filter and transition zones (of well-
compacted finer sized rockfill) on the downstream side.

For a number of case studies the lateral displacement ratio versus fill height plot
(Figure 7.13) shows a marked increase in the rate of lateral displacement ratio at fill
heights in the range 20 to 50 m above the gauge. The cases that display this behaviour
are typically clayey earthfills placed at or wet of Standard optimum. It is possible that
this response may be due to yielding or may reflect an increase in the undrained
Poisson’s ratio as the initially over-consolidated earthfill approaches or becomes
normally consolidated.
For several case studies the lateral displacement ratio of the core was compared to
the pore water pressure response at similar elevations to ascertain if the incremental
increase in lateral displacement ratio observed correlated to a change in the pore water
pressure response (Figure 7.14). The results indicate:
• At Thomson dam (h/H = 0.53) and Talbingo dam (h/H = 0.59) the data indicates that
the increase in incremental lateral displacement ratio occurs at about the same level
Chapter 7: Deformation behaviour of embankment dams Page 7.49

of total vertical stress level as an incremental increase in the pore water pressure
response.
• At Dartmouth dam (Figure 7.14a) there is some indication of an incremental
increase in the lateral displacement ratio and the pore water pressure response at 20
to 25 m fill height, although it is not clearly evident for the lateral displacement
ratio.
• At Blowering dam the pore water pressure response at Piezometer 16, at a lower
elevation than the deformation gauge, shows a similar response at 22 m fill height,
but the piezometer at the same elevation (Piezometer 23) shows no incremental
increase in the pore water pressure response.

Overall, there is some data, although it is not conclusive, that the observed
incremental increase in lateral displacement ratio for several of the wet placed clayey
earthfill case studies occurs concurrently with an incremental increase in the pore water
pressure response.
Several case studies show a flattening in the curve of lateral displacement ratio
versus fill height near the end of construction (Figure 7.13). This is potentially due to
arching (El Infiernillo) and the effect of embankment shape on the distribution of
vertical and lateral stresses, which is not properly allowed for by considering depth of
fill as the variable.

a) -3.0% 1500
IHM1-A1 at EL 370m (h/H = 0.71)
Pore water pressure (kPa) .

-2.5% 1250
Lateral displacement ratio (%)

Piezo DPM31 (EL 370 m)

-2.0% 1000

-1.5% 750
.

-1.0% 500
Dartmouth Dam:
180m high dam with crest at EL 494 m.
-0.5% Earthfill core of low to medium plasticity clayey 250
to silty sand placed at close to OMC.

0.0% 0
0 20 40 60 80 100 120 140
Fill height above gauge (m)
Chapter 7: Deformation behaviour of embankment dams Page 7.50

b) -2.5% 750
IHM1-A1 at EL 316 m (h/H = 0.65)

Pore water pressure (kPa) .


Lateral displacement ratio (%)
-2.0% Piezo HP23 at EL 313 m (h/H = 0.68) 600
Piezo HP16 at EL 295 m (h/H = 0.84)

-1.5% 450

-1.0% 300
.

-0.5% 150
Blowering Dam:
112m high dam with crest at EL 389 m.
0.0% Low to medium plasticity clayey sand to sandy clay
0
core placed 0.3% wet of Standard optimum.
0.5% -150
0 10 20 30 40 50 60 70 80 90 100
Fill height above gauge (m)

c) -3.0% 800
LDR at EL 457m (h/H = 0.59)

Pore water pressure (kPa).


Piezo HP23 at EL 457 m (h/H = 0.59)
Lateral displacement ratio

600
-2.0% Talbingo Dam:
162m high dam with crest at EL 552 m.
Earthfill core of low to medium plasticity gravelly
(%) .

sandy clay placed at close to OMC. 400

-1.0%
200

0.0% 0
0 250 500 750 1000 1250 1500 1750 2000
Measured total vertical stress from pressure cell PC23 (kPa)

d) -3.5% 700
IHM1-13 at EL. 344 m (h/H = 0.71)
-3.0% IHM2-12 at EL. 374 m (h/H = 0.53) 600

Pore water pressure (kPa) .


Lateral displacement ratio

Piezo 27 at EL 344 m
-2.5% Piezo 40 (EL 374 m) 500

Thomson Dam: 166m high, crest EL 462 m.


-2.0% 400
(%) .

medium plasticity clayey sand core placed 2%


wet of OMC.
-1.5% 300

-1.0% 200

-0.5% 100

0.0% 0
0 200 400 600 800 1000 1200 1400 1600
Measured total vertical stress from pressure cells (kPa)

Figure 7.14: Lateral displacement ratio and pore water pressure versus fill height or
measured total vertical stress at (a) Dartmouth dam; (b) Blowering dam; (c) Talbingo
dam and (d) Thomson dam.
Chapter 7: Deformation behaviour of embankment dams Page 7.51

7.4.3 VERTICAL DEFORMATION OF THE CORE DURING CONSTRUCTION

Internal settlements during construction were collected for 57 embankments of earthfill,


zoned earthfill and zoned earth and rockfill types. The settlement records were mostly
from internal vertical settlement gauges (IVMs) installed in the central core of the
earthfill (near the dam axis) as construction proceeded.
The purpose of the analysis is to define “normal” deformation behaviour so case
studies showing “abnormal” deformation can be identified and then further analysed.
The data is grouped into those case studies where a “homogeneous” type analysis (i.e.
small lateral core strain) is appropriate, and those where lateral deformations are likely
to be significant.
The vertical deformation in the core under the embankment axis (or within close
proximity to the axis) is presented in the following formats:
• Vertical strain versus effective vertical stress (and secant modulus versus effective
vertical stress) for those case studies where a “homogeneous” type analysis is
appropriate.
• Vertical strain versus fill height above the cross-arm for case studies where lateral
deformations are likely to be significant.
• Total vertical settlement (in millimetres) of the core during the period of
construction versus embankment height. The total vertical settlements are estimated
from the cumulative settlement between the cross-arm gauges.

Vertical strains in the core are calculated between the individual cross-arm intervals
of the IVM by dividing the measured settlement (between the cross-arm interval) by the
original distance between the cross-arms. The vertical stresses or fill heights are
representative of the midpoint between the nominated cross-arm intervals. Corrections
have been made to the strain readings so that the zero strain reading is equivalent to zero
stress (or zero fill height). The corrections are based on the strain measurements at low
stress levels.
The vertical strain (and secant modulus) plots are presented in two forms; (i) vertical
strain at the end of construction, and (ii) vertical strain during construction. For the
plots at the end of construction the vertical strains (or secant moduli) are calculated and
plotted at cross-arm intervals over the full depth of the IVM gauge from near to crest
level to foundation level. These plots, which include all the case studies, are referred to
as the “vertical strain profile at end of construction” and they broadly define the range
Chapter 7: Deformation behaviour of embankment dams Page 7.52

of the data. The deformations in the upper 10 to 20% of the embankment have
generally been excluded because of narrowing of the embankment and reduction in the
lateral confinement, which often gives relatively large vertical strains resulting in low
estimates of apparent moduli.
The vertical strain (and secant modulus) plots during construction (as opposed to the
plots at the end of construction) only present data from selected cross-arm intervals
from selected representative case studies. For these cases the cross-arm interval and its
h/H ratio are identified in the legend. The purpose of these plots is to highlight the
change in vertical strain (or secant modulus) with increasing vertical stress (or fill
height), which is more clearly evident in these plot types.

7.4.3.1 Vertical Deformation for Dry Placed Earthfills and Other Cases with Small
Lateral Core Displacement
The types of case studies analysed in this section are those for which simplification to
an essentially one-dimensional analysis of the earthfill core is considered a suitable
approximation of the deformation behaviour, and they are:
• Embankments with relatively broad core widths and dry placed earthfill cores,
typically those placed at moistures contents drier than about 0.5 to 1% dry of
Standard optimum. The limitation on core width is for case studies with thick cores
(combined slope greater than 1H to 1V), but several case studies with broader
medium width cores have been included.
• Embankments with cores of dominantly sandy to gravelly soil types with low
plasticity silty fines or with low plasticity clayey fines of low fines content, and
within which positive pore water pressures during construction were relatively low.
The limitation on pore water pressure development generally restricts these cases to
medium to thick core widths that are generally placed on the dry side of Standard
optimum moisture content.
• Zoned embankments with thin cores where the compressibility properties of the core
are similar to the shoulders. Two cases have been included Naramata and Frauenau,
both dominantly sandy and gravelly cores with silty fines placed on the dry side of
optimum.
Chapter 7: Deformation behaviour of embankment dams Page 7.53

Figure 7.15 presents vertical strain versus vertical effective stress as the
embankment is raised for selected cross-arm intervals of selected case studies. The data
is from the lower third of the core (i.e. h/H > 0.66). Effective vertical stresses at the
midpoint between the nominated cross-arms have been estimated from total stresses
estimated as described in Section 7.4.1.1 with positive pore water pressures then
deducted. Note that negative pore water pressures at low stresses have been ignored
and therefore the actual effective stresses at these low stress levels will be higher than
indicated in the Figure 7.15 and other figures in this section.
The plots (Figure 7.15) show an apparent pre-consolidation pressure beyond which
the incremental vertical strain (in terms of the log of stress) is greater. From the
intersection of the two parts of the curves, these apparent pre-consolidation pressures
range from 700 to 1000 kPa for the silty gravel, silty sand and clayey sand (with low
content fines) earthfills, to 200 to 400 kPa for the clayey earthfills. The influence of
negative pore water pressures, which has been ignored, will steepen the stress-strain
relationship at low stress levels in these partially saturated earthfills.
Figure 7.16 presents the secant modulus versus vertical effective stress as the
embankment is raised for selected cross-arm intervals of selected case studies; the
modulus calculated by dividing the estimated vertical effective stress by the vertical
strain. The modulus has been termed a “confined secant modulus” based on the
assumption of negligible lateral strain. Figure 7.16 shows a relatively broad variation in
confined secant modulus at a given stress level, more so for the dominantly sandy and
gravelly core types probably because of the effect of variation in fines content and fines
plasticity as well as compaction conditions.
For most cases shown in Figure 7.16 the confined secant modulus generally
increases with increasing effective vertical stress, but with a large variation in the
incremental change in modulus over a given stress level. Other cases show the confined
secant modulus to remain relatively steady with increasing stress whilst some even
show it to decrease. A gradual increase in the confined secant modulus with increasing
stress, for the stress range of the data, is considered the typical response indicating the
earthfill becomes less compressible as the voids ratio decreases.
Chapter 7: Deformation behaviour of embankment dams Page 7.54

a) 0%

1%

Vertical strain (%)


2%

3%

Chaffey Dam - IVM 1, Xarms 3 to 4 (h/H = 0.85)


4% Burrendong Dam - IVM B, Xarms 4 to 6 (h/H = 0.92)
Horsetooth Dam - IVM A, Xarms 5 to 7 (h/H = 0.80)
Navajo Dam - IVM A, Xarms 3 to 5 (h/H = 0.92)
5% Steinaker Dam - IVM A, Xarms 3 to 4 (h/H = 0.83)
Upper Yarra Dam - ES1, Xarms 5 to 8 (h/H = 0.91)
Burrendong Dam - IVM A, Xarms 8 to 10 (h/H = 0.84)
6%
10 100 1000 10000
Estimated effective vertical stress (kPa)

b) 0%

1%
Vertical strain (%)

2%

3%
Frauenau Dam - X-arm 3 to 4 (h/H = 0.67)
Copeton Dam - IVM A, Xarms 11 to 13 (h/H = 0.85)
4% Wyangala Dam - IVM A, Xarms 3 to 5 (h/H = 0.94)
Bradbury Dam - IVM A, Xarms 10 to 12 (h/H = 0.78)
5% Burrendong Dam - IVM C, Xarms 5 to 7 (h/H = 0.90)
Trinity dam - IVM B, Xarms 2 to 6 (h/H = 0.94)

6%
10 100 1000 10000
Estimated effective vertical stress (kPa)
Figure 7.15: Vertical stress versus strain of the core during construction for selected
cross-arm intervals of selected case studies; (a) dry placed clay cores, and (b) dry placed
dominantly sandy and gravelly cores with plastic fines.

a) 70
Chaffey Dam - IVM 1, Xarms 3 to 4 (h/H = 0.85)
Chaffey Dam - IVM 1, Xarms 7 to 8 (h/H = 0.68)
Confined secant moduli (MPa) .

60 Burrendong Dam - IVM B, Xarms 4 to 6 (h/H = 0.92)


Burrendong Dam - IVM A, Xarms 8 to 10 (h/H = 0.84)
Horsetooth Dam - IVM A, Xarms 5 to 7 (h/H = 0.80)
50 Navajo Dam - IVM A, Xarms 3 to 5 (h/H = 0.92)
Upper Yarra Dam - ES1, Xarms 5 to 8 (h/H = 0.91)
Steinaker Dam - IVM A, Xarms 3 to 4 (h/H = 0.83)
40

30

20

10

0
0 250 500 750 1000 1250 1500 1750 2000
Estimated effective vertical stress (kPa)
Chapter 7: Deformation behaviour of embankment dams Page 7.55

b) 80

Confined secant moduli (MPa) .


70

60

50

40
Copeton Dam - IVM A, Xarms 11 to 13 (h/H = 0.85)
30 Wyangala Dam - IVM A, Xarms 3 to 5 (h/H = 0.94)
Bradbury Dam - IVM A, Xarms 2 to 4 (h/H = 0.96)
20 Burrendong Dam - IVM C, Xarms 5 to 7 (h/H = 0.90)
Bradbury Dam - IVM A, Xarms 10 to 12 (h/H = 0.78)
Trinity dam - IVM B, Xarms 2 to 6 (h/H = 0.94)
10 Frauenau Dam - X-arm 3 to 4 (h/H = 0.67)
Trinity dam - IVM B, Xarms 14 to 18 (h/H = 0.67)
0
0 500 1000 1500 2000 2500 3000
Estimated effective vertical stress (kPa)

Figure 7.16: Confined secant moduli of the core versus vertical stress during
construction for selected cross-arm intervals of selected case studies, (a) dry placed clay
cores, and (b) dry placed dominantly sandy and gravelly cores with plastic fines.

Of the case studies that show an increase in the confined secant modulus with increasing
stress, the secant modulus at low stress levels is sometimes quite low; for example, the
cross-arm intervals shown for Burrendong, Navajo and Steinaker dams. Evaluation of
the stress-strain relationship for several of these cases (Figure 7.17) shows that the
tangent confined moduli at low stresses is much lower than that at stress levels in excess
of 200 to 300 kPa, and is near constant at the higher stress levels. Given that the
earthfill core for these embankments has been placed at moisture contents on the dry
side of Standard optimum and well-compacted in thin lifts, this type of stress-strain
behaviour is unlikely. It would be expected that the tangent confined moduli at low
stresses would be at least similar to that observed at stresses up to 500 to 1000 kPa.
The most likely cause of this observation is suspected to be settlement on seating of
the cross-arm after installation possibly due to a slightly uneven bedding surface or
loose nature of the bedding material on which the cross-arm sits. Settlements of 20 to
30 mm due to seating can make a large difference in the calculated tangent and confined
secant moduli at low stresses.
Chapter 7: Deformation behaviour of embankment dams Page 7.56

2000

Estimated effective vertical stress (kPa) . E tang = 40 MPa

1500 E tang = 65 MPa

E tang = 40 MPa
1000

seating of cross-arm E tang = 34 MPa


at low stresses

500 Navajo Dam - IVM A, Xarms 3 to 5 (h/H = 0.92)

Navajo Dam - IVM A, Xarms 10 to 12 (h/H = 0.75)

Steinaker Dam - IVM A, Xarms 3 to 4 (h/H = 0.83)


0 Wyangala Dam - IVM A, Xarms 3 to 5 (h/H =
0% 1% 2% 3% 4% 5% 6%
Vertical Strain (%)

Figure 7.17: Seating settlements at low stresses suspected as cause of low moduli
estimates at low stresses.

Table 7.13 presents a summary of the range and average of the confined secant moduli
at various effective vertical stress levels from the case studies for well-compacted dry
placed central core zones. It includes:
• Data from the case studies represented in Figure 7.15, including values from
analyses of other cross-arm intervals in these embankments that are not shown.
• Data from case studies where the vertical strain was also analysed for specific cross-
arm intervals but not shown in Figure 7.15.
• Data from the end of construction vertical strain profiles (Figure 7.18 and Figure
7.19).
• Data from Gould (1953, 1954) for a number of USBR dams of dry placed earthfills
mainly from embankments with very broad earthfill zones. Gould presented his
data in terms of vertical strain versus effective vertical stress. It has been converted
here to values of secant moduli.

Seating effects have only been considered in the confined secant modulus values for
those cases where the stress-strain relationship indicated this to have had a clearly
identifiable influence on the deformation behaviour. For most of the cases represented
in Table 7.13 the data was not available from which to make an assessment of seating.
Therefore, the real lower limits and averages, particularly at the lower vertical stress
levels, are possibly slightly higher than indicated. Categorisation based on material type
Chapter 7: Deformation behaviour of embankment dams Page 7.57

is justified from the findings by Gould (1954) and from analysis of the case study data
collected for this study.

Table 7.13: Confined secant moduli during construction for well-compacted, dry placed
earthfills.
Confined Secant Modulus (MPa)*1
No. Effective Vertical Stress Range (kPa)
Core Material Type
Cases 500 to 700*2 1000 1500 to 2000
Clayey soil types (CL) – sandy clays 20 15 to 45 15 to 50 25 to 42
and gravelly clays (27) (30)
Silty soils (ML), data from Gould. 3 30 to 60 - -
Clayey Sands to Clayey Gravels 11 20 to 65 30 to 65 25 to 65
(SC/GC), plastic fines, > 20% fines (33) (40)
Silty Sands to Silty Gravels 9 35 to 65 45 to 65 35 to 65
(SM/GM), plastic fines, > 20% fines (46) (50)
Sandy and Gravelly soils - non- 6 35 to 80 35 to 90 40 to 90
plastic fines or < 20% plastic fines (60) (60)
Notes *1 Values represent range of confined secant moduli, values in brackets represent average
*2 Includes Gould (1953, 1954) data.

The range of confined secant moduli within any core material type is relatively broad,
up to 0.5 to 2 times the mean. Part of this variation is likely explainable by variations in
material properties (particle size distribution, plasticity, etc.), variations in compacted
density ratio and inaccuracies in the IVM measurement readings. Correlation between
the compacted density ratio at placement (and dry density) and confined secant moduli
for a given earthfill type did not identify any clear trend, however, the database was
limited.
Figure 7.18 and Figure 7.19 present the vertical strain and confined secant moduli
versus effective vertical stress at end of construction for all case studies of dry placed
earthfills. As discussed earlier, they differ from Figure 7.15 and Figure 7.16, which are
for selected cross-arm intervals of selected case studies during construction. Essentially
there is no difference between the plots, but Figure 7.18 and Figure 7.19 present the data
for all case studies representative of dry placed earthfills and show a broader range of
variation of vertical strain and modulus.
Chapter 7: Deformation behaviour of embankment dams Page 7.58

a) 0%

1%
Vertical strain in core (%) .
2%

3%

4%

5%

6%

7%

8%
0 250 500 750 1000 1250 1500 1750 2000
Estimated effective vertical stress (kPa)
Windemere - ES1 Chaffey - IVM A Glenbawn Saddle - IVM A Deer Creek
Medicine Creek - IVM A Navajo - IVM A Steinaker - IVM A Upper Yarra - ES1
Upper Yarra - ES2 Eppalock - IVM 1 Sasumua Glenbawn - IVM A
Burrendong - Zone 1A Burrendong - Zone 1C Horsetooth - IVM A Horsetooth - IVM B
San Luis - IVM D San Luis - IVM B

b) 0%
Likely arching across
1% narrow core at Naramata
Vertical strain in core (%)

2%

3%

4%

5%

6% approximate bounds for sandy and


At Hirakud - strains up to gravelly earthfills with plastic fines
13.5% measured at stresses
7% up to 860 kPa.

8%
0 500 1000 1500 2000 2500 3000
Estimated effective vertical stress (kPa)
Naramata Frauenau Andong Bradbury - IVM A Copeton - IVM A
Copeton - IVM B Copeton - IVM C Wyangala - IVM A Wyangala - IVM B Burrendong - Zone 1B
Granby - IVM A Granby - IVM B Deer Creek Hirakud Trinity - IVM A

Figure 7.18: Vertical strain versus effective vertical stress in the core at end of
construction for (a) dry placed clay earthfills and (b) dry placed dominantly sandy and
gravelly earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.59

a) 80

70
Confined secant moduli (MPa) .
60

50

40

30

20

10

0
0 250 500 750 1000 1250 1500 1750 2000
Estimated effective vertical stress (kPa)
Deer Creek Chaffey - IVM A Windemere - ES1 Medicine Creek - IVM A
Glenbawn Saddle - IVM A Navajo - IVM A Steinaker - IVM A Upper Yarra - ES1
Upper Yarra - ES2 Eppalock - IVM 1 Sasumua Glenbawn - IVM A
Burrendong - Zone 1A Burrendong - Zone 1C Horsetooth - IVM A Horsetooth - IVM B
San Luis - IVM D San Luis - IVM B

b) 120
Confined secant modulus (MPa) .

100

80

Likely arching across


narrow core at Naramata
60

40

20
abnormally low moduli at Hirakud
dam (refer Section 6.2.2)
0
0 500 1000 1500 2000 2500 3000
Estimated effective vertical stress (kPa)
Naramata Frauenau Andong Bradbury - IVM A Copeton - IVM A
Copeton - IVM B Copeton - IVM C Wyangala - IVM A Wyangala - IVM B Burrendong - Zone 1B
Granby - IVM A Granby - IVM B Deer Creek Hirakud Trinity - IVM A

Figure 7.19: Confined secant moduli versus effective vertical stress in the core at end of
construction for (a) dry placed clay earthfills and (b) dry placed dominantly sandy and
gravelly earthfills.

For the dominantly sandy and gravelly earthfill types the data (Figure 7.18b and Figure
7.19b) show:
Chapter 7: Deformation behaviour of embankment dams Page 7.60

• Higher confined secant moduli for silty sand and silty gravel type earthfills with
non-plastic fines, or with low plastic fines content (Bradbury, Naramata, Andong
and Frauenau dams).
• On average, lower confined secant moduli in the clayey sand and clayey gravel type
earthfills (Copeton, Granby, Deer Creek and Burrendong (Zone 1B) dams).
• Likely arching effects in the narrow, dry placed earthfill core at Naramata.

For the clay earthfills the plots (Figure 7.18a and Figure 7.19a) show a reasonable
degree of consistency between case studies. Several of the points that show high strains
at a given vertical stress are within wetter placed earthfills near to the interface with the
foundation. In general, relatively high vertical strains are observed at this interface
region, most probably due to plastic type deformations of the wet placed earthfill in the
foundation contact zone.

7.4.3.2 Vertical Deformation for Wet Placed Earthfills and Other Cases with Large
Lateral Core Displacement

The types of case studies analysed in this section are zoned embankments where the
deformation of the earthfill core cannot be adequately modelled by simplification to a
one-dimensional analysis. For these case studies the lateral displacement of the core
during construction is generally large and has a significant influence on the vertical
deformation behaviour of the core. Analysis of the deformation behaviour of the core
should be undertaken on a case-by-case basis using finite element methods to model the
embankment construction. Notwithstanding this, there is some merit in comparative
analysis of the case studies as a general guide to possible identification of “abnormal”
deformation behaviour.
The figures in this section are presented versus fill height. It would have been
preferable to use total vertical stress, but the data includes a number of case studies
where arching was significant and no information of total stresses from pressure cells
was available. Rather than attempt to estimate the total vertical stress the fill height was
used for all case studies. The case study data on the vertical strain in the core has been
sorted based on core width and is presented in the following figures:
• Figure 7.20 – incremental vertical strain versus fill height for selected cross-arm
intervals from selected case studies.
Chapter 7: Deformation behaviour of embankment dams Page 7.61

• Figure 7.21 and Figure 7.22 – vertical strain profile of the core (i.e. versus fill
height) at end of construction. Figure 7.21 is of dominantly sandy and gravelly
earthfill cores and Figure 7.22 is of clay cores.

The difference between the plot types, as previously discussed, is that the
incremental vertical strain plots highlight the change in vertical strain with increasing
fill height, and the vertical strain profiles at end of construction include all the case
studies and highlight the degree of variation from the data.
From Figure 7.20, the general trend for thin cores is for an apparent decrease in the
incremental vertical strain with increase in fill height. The medium width cores show a
similar trend but it is not as evident as it is for the thin cores. Arching across the core is
likely to be a significant factor influencing this behaviour, particularly for the thin cores.
If measured total vertical stress were used as the x-axis this trend would probably not be
evident for most of the case studies because the influence of arching would be included
in the total vertical stress. The analysis of the core deformation during construction at
Blowering dam shows this to be the case (refer Section 6 of Appendix G).
Figure 7.21 and Figure 7.22 highlight the very high vertical strains measured within
some regions of the core of several zoned embankments including Blowering, Beliche,
Tedorigawa and Chicoasen dams. High lateral strains and plastic deformation of the
wet placed earthfill core are likely to be the reason for the large vertical strains observed
in most, but not all cases. Further discussion on several of these cases is presented in
Section 7.9.
In Figure 7.21 and Figure 7.22 the bounds shown on several of the figures are an
interpretation, based on evaluation of the data, of the general limits of “normal” type
deformation during construction in wet placed earthfills. However, assessment of
whether or not a point or series of points located outside the bounds shown is potentially
“abnormal” requires further evaluation, and this is the main purpose of the indicative
bounds. In general, for those cases where a series of points plot below the lower bound
(such as for Beliche dam in Figure 7.21a) or one or two points plot well below the lower
bound it is likely that the deformation behaviour is an outlier to the general trend.
Chapter 7: Deformation behaviour of embankment dams Page 7.62

a) 0%

1%
Vertical strain between x-arms (%) .
2%

3%

4%

5%

6%

7%

8%

9%
0 20 40 60 80 100 120 140 160 180
Fill height above the midpoint between the cross-arm interval (m)
Chicoasen Dam - IVM IB-4, Xarms 5 to 6 (h/H = 0.59) Beliche Dam - IVM ES1, Xarms 10 to 11 (h/H = 0.48)
El Infiernillo Dam - IVM I1, Xarms 5 to 6 (h/H = 0.82) Beliche Dam - IVM ES1, Xarms 5 to 6 (h/H = 0.75)
El Infiernillo Dam - IVM I1, Xarms 17 to 18 (h/H = 0.70) Tedorigawa Dam - Xarms a to base (h/H = 0.91)
Kremasta Dam - IVM S1, Xarms 5 to 6 (h/H = 0.84) Tedorigawa Dam - Xarms a to b (h/H = 0.75)
Kremasta Dam - IVM S1, Xarms 11 to 12 (h/H = 0.64) Tedorigawa Dam - Xarms c to d (h/H = 0.47)
Kastraki Dam - IVM S1, Xarms 9 to 10 (h/H = 0.70)

b) 0%
Vertical strain between x-arms (%) .

1%

2%

3%

4%

5%

6%

7%

8%

9%
0 20 40 60 80 100 120 140 160 180
Fill height above the midpoint between the cross-arm interval (m)
Dartmouth Dam - IVM ES1, Xarms 2 to 4 (h/H = 0.96) Dartmouth Dam - IVM ES1, Xarms 9 to 11 (h/H = 0.80)
Dartmouth Dam - IVM ES1, Xarms 30 to 32 (h/H = 0.36) Talbingo Dam - IVM ES1, Xarms 1 to 3 (h/H = 0.98)
Talbingo Dam - IVM ES1, Xarms 9 to 12 (h/H = 0.82) Talbingo Dam - IVM ES1, Xarms 19 to 22 (h/H = 0.62)
Blowering Dam - IVM A, X-arms 4 to 5 (h/H = 0.90) Blowering Dam - IVM A, X-arms 9 to 12 (h/H =0.74)
Blowering Dam - IVM A, X-arms 15 to 17 (h/H = 0.59) Blowering Dam - IVM A, X-arms 22 to 24 (h/H = 0.40)

Figure 7.20: Vertical strain in the core versus fill height during construction at selected
cross-arm intervals in selected case studies of earthfills placed close to or wet of
Standard optimum, for (a) thin core (combined slopes less than 0.5H to 1V), and (b)
medium cores (combined slope ≥ 0.5H to 1V and < 1H to 1V).
Chapter 7: Deformation behaviour of embankment dams Page 7.63

a) 0%
Chicoasen - IVM I-B4
1% Sagae
Vertical strain in core (%) . Tedorigawa
2% Beliche - IVM A

3%

4%

5%

6%

7% large plastic deformations of


the core
8%
0 25 50 75 100 125 150 175 200
Depth below crest to the midpoint between cross-arms (m)

b) 0%

1%

indicative bounds of "normal"


Vertical strain in core (%)

2% type deformation

3%

4%

5%

6%
more likely isolated points due to
7% inaccuracies in measurement
rather than outliers

8%
0 25 50 75 100 125 150 175 200
Depth below crest to the midpoint between cross-arms (m)
Fukada - IVM H2 Shirakawa Tooma - IVM A
Soyang Dartmouth - IVM ES2 Dartmouth - IVM ES1
Parangana - IVM B Parangana - IVM C Mammoth Pool - IVM C4

Figure 7.21: Vertical strain in the core versus fill height at end of construction for
dominantly sandy and gravelly earthfills with plastic fines placed close to or wet of
Standard optimum, for (a) thin cores (combined slopes less than 0.5H to 1V), and (b)
medium to thick cores (combined slope ≥ 0.5H to 1V).
Chapter 7: Deformation behaviour of embankment dams Page 7.64

a) 0%
El Infiernillo - IVM I1
1% Netzahualcoyotl - IVM I3
Vertical strain in the core (%) . Wivenhoe - IVM ES1
Wivenhoe - IVM ES4
2%
Kastraki - IVM S1
Kastraki - IVM S2
3% Kremasta - IVM S1
Kremasta - IVM S2
4%

5%

6%

7%

8%
0 20 40 60 80 100 120 140 160
Depth below crest to the midpoint betweeen cross-arms (m)

b) 0%
Vertical strain in the core (%) .

2%

4%
Talbingo dam - low
strains due to arching
6%

Blowering - IVM A (wet)


8% Blowering - IVM A (0.2% dry)
Talbingo - IVM 1
Talbingo - IVM 2
10% Maroon - IVM ES1
indicative bounds of
Maroon - IVM ES2
"normal" type deformation
12% Maroon - IVM ES3
Agigawa
Split Yard Creek - IVM 1
14%
0 20 40 60 80 100 120 140 160
Depth below crest to the midpoint between cross-arms (m)

Figure 7.22: Vertical strain of the core versus fill height at end of construction for clay
earthfills placed close to or wet of Standard optimum moisture content, for (a) thin core
(combined slopes less than 0.5H to 1V), and (b) medium to thick cores (combined slope
≥ 0.5H to 1V).

For several case studies the point or points that plot outside the bounds are not
necessarily indicative of an outlier or potentially “abnormal” deformation behaviour.
Several case studies show a large degree of variation in vertical strain between
consecutive cross-arm or depth intervals, showing an almost zigzag type pattern (e.g.
IVM ES1 at Dartmouth in Figure 7.21b). This type of effect is, in some cases, due to
the inaccuracies or errors of measurement in the IVM gauge and does not necessarily
represent a large localised region of vertical strain. These points are not likely to be
outliers.
Chapter 7: Deformation behaviour of embankment dams Page 7.65

Conversely, it should also be recognised that a case study that shows potentially
“abnormal” type deformation behaviour in the core during construction may not
necessarily be identified as such in plots such as Figure 7.21 and Figure 7.22.
Although, of the case studies represented in the plots those within the bounds are all
considered to be indicative of “normal” type deformation behaviour.

7.4.3.3 Total Settlement of the Core During the Construction Period


Another type of plot considered useful in the identification of potential “abnormal”
deformation behaviour of the core during construction is the total settlement of the core
that occurs during the period of embankment construction. Figure 7.23 presents the
total settlement of the earthfill core or central earthfill zone during the period of
construction versus embankment height for the case studies. The results show a general
trend of increasing settlement (as a percentage of the embankment height) with
increasing embankment height. At embankment heights of less than 50 m, total
settlements were generally in the order of less than 1% up to 2.5%, and for embankment
heights greater than 150 m, in the order of 2 to 5%.
Several sub-groups of material type and core width with a similar total settlement
versus embankment height relationship are identifiable (Figure 7.23b), they are:
• Dominantly fine grained earthfills, mainly clays but includes some dominantly silty
soils, covering thin to very broad cores and dry to wet placement conditions. These
cases generally plot on the higher side with settlements in the order of 3 to 3.5% for
embankment heights greater than about 100 m.
• Dominantly sandy and gravelly earthfills with plastic fines (i.e. clayey and silty
sands and gravels), dry to wet placement conditions and medium width to very
broad core sizes. The total settlement for this sub-group is similar to that for the
clay earthfills, reaching settlements in the order of 3 to 4% for embankment heights
greater than about 100 m.
• Dominantly sandy and gravelly earthfills with non-plastic fines or with low content
plastic fines (typically less than 10 to 20%) for thin to thick core widths. The total
settlement for these core types is at the low end for embankment dams,
approximately half that of the clayey earthfill types at a given embankment height.
The low settlement reflects the low compressibility of these earthfills when well
Chapter 7: Deformation behaviour of embankment dams Page 7.66

compacted. Settlements are typically in the range 1 to 2% for embankment heights


up to 150 m.
• Dominantly sandy and gravelly earthfills with thin width cores. The total settlement
for these core types is in between that of the clayey earthfills and non-plastic sandy
and gravelly earthfills, reaching maximum settlements in the range 1.5 to 3% for
embankment heights greater than about 100 m.

The data within each material / core width sub-group shows a reasonable fit as
indicated by the high regression coefficients of the trendlines (Figure 7.23b). The
trendlines are quadratic equations and are given in Table 7.14 along with the standard
errors and regression coefficients. Outliers to the general trend (Beliche, Hirakud,
Blowering, Tedorigawa and Nurek dams) were excluded from the correlations. These
cases are discussed further in Section 7.9.
The settlement to embankment height curves from Figure 7.23b are useful as a
guideline for estimation of the total core settlement during construction and
identification of potentially “abnormal” deformation behaviour for embankment heights
up to 100 to 125 m. Above about 125 m only one or two points define the shape of the
relationships and therefore settlement estimates will be less reliable.
Correlations between the settlement as a percentage of the embankment height and
embankment height in the form of a power function are summarised in Table 7.15. The
regression coefficients indicate a good correlation exists for the sub groups and the
range in standard error is relatively low, from 0.24 to 0.47%.
The correlation shown for several of the sub groups, in particular for the clayey
earthfills, is somewhat coincidental. The clay earthfill sub-group includes dry to wet
placement conditions and core widths from thin to very broad. It is evident from
previously presented data that the deformation behaviour of a wet placed, thin to
medium width clayey core is significantly different to that of a dry placed, very broad
clayey earthfill core. Therefore, for these two different material placement conditions
and core width geometries to be included in the same sub group is somewhat
coincidental.
Chapter 7: Deformation behaviour of embankment dams Page 7.67

a)
14000
Clayey and silty cores
5%
Sandy and gravelly cores - medium to thick, plastic fines 6% Nurek
Sandy and gravelly cores - non plastic or low fines content
12000
Sandy and gravelly cores - thin
.

4%
Core settlement during construction (mm)

10000

3%
8000

Blowering
6000
2%

4000

Beliche
1%
Hirakud
2000

0
0 50 100 150 200 250 300
Embankment height (m)

b) 8000
4%
Sandy and Gravelly Cores
(GM/GC/SM/SC) - medium to
7000
thick cores, plastic fines
Core settlement during construction (mm)

2
R = 0.97

6000 Tedorigawa
Clayey and Silty Cores Blowering
(CL/CH/ML/MH)
2
5000 R = 0.96

6%

4000 2%

3000 Sandy and Gravelly


Beliche Cores (GM/GC/SM/SC) -
Hirakud
thin cores
2
2000 R = 0.94

Sandy and Gravelly Cores


1000 (GM/SM/GP/SP) - non
plastic or low fines content
2
R = 0.98
0
0 50 100 150 200
Embankment height (m)
Clayey and silty cores Sandy and gravelly cores - medium to thick, plastic fines
Sandy and gravelly cores - non plastic or low fines content Sandy and gravelly cores - thin

Figure 7.23: Core settlement during construction of earth and earth-rockfill


embankments (a) including Nurek dam, and (b) excluding Nurek dam.
Chapter 7: Deformation behaviour of embankment dams Page 7.68

Table 7.14: Equations of best fit for core settlement versus embankment height during
construction
Core Material / Shape No. Equation for Core R2 *2 Std. Err.
Type Cases Settlement*1 (mm) *3
Clay cores – all sizes 42 Settlement = H (0 .152 H + 12.60 ) 0.96 275

Sandy and gravelly cores:


- medium to thick, plastic 25 Settlement = H (0.183 H + 7 .461) 0.97 290
- thin 5 Settlement = H (0.136 H + 2.620 ) 0.94 635
- non-plastic 7 Settlement = H (0.063 H + 8 .57 ) 0.98 130
Note: *1 settlement in millimetres, embankment height H in metres.
*2 R2 = regression coefficient
*3 Std. Err. = standard error of the settlement

Table 7.15: Equations of best fit for core settlement (as a percentage of embankment
height) versus embankment height during construction
Core Material / Shape No. Equation for Core R2 *2 Std. Err.
Type Cases Settlement*1 (%) *3
Clay cores – all sizes 42 Settlement (% ) = 0.179 H 0.60 0.77 0.41

Sandy and gravelly cores:


- medium to thick, plastic 25 Settlement (% ) = 0.079 H 0.76 0.81 0.39

- thin 5 Settlement (% ) = 0.063 H 0.72 0.70 0.47

- non-plastic 7 Settlement (% ) = 0.187 H 0. 46 0.71 0.24

Note: *1 settlement as a percentage of the embankment height, embankment height H in metres.


*2 R2 = regression coefficient
*3 Std. Err. = standard error of the settlement (as a percent of embankment height)
Chapter 7: Deformation behaviour of embankment dams Page 7.69

7.5 GENERAL DEFORMATION B EHAVIOUR ON F IRST F ILLING OF


EARTHFILL AND ZONED EARTH AND EARTH -ROCKFILL
EMBANKMENTS
The following discussion summarises the factors affecting the deformation behaviour of
earthfill and earth – rockfill embankments during first filling of the reservoir, drawing
on relevant information from the published literature. The discussion does not go into
case study specific details, but some of the issues raised are discussed in greater detail in
the individual case study analysis in Sections 7.9 to 7.11.
Analyses of first filling generally consider the main water barrier element to be
impermeable during first filling and the water load to act on the upstream face of this
zone. Permeable earth and rockfill zones upstream of the main earthfill zone will
become saturated during first filling. Therefore, the factors affecting the deformation
behaviour of embankments during reservoir filling are primarily the water load, the
compressibility properties of the earthfill core and downstream shoulder, and the
susceptibility to collapse compression on wetting of the permeable earth and rockfill
zones upstream of the core. Other factors that can affect the embankment deformation
during first filling include time dependent deformations, such as creep and earthfill
consolidation, and differential settlement of the foundation, neither of which is
considered in detail in this section.
Nobari and Duncan (1972b), Sherard (1973) and others, describe the effect of the
water load (Figure 7.24) as threefold:
(i) Applied water load on the upstream side of the core or earthfill forming the water
barrier, resulting in a net increase in the total stresses within and downstream of
this zone. The resultant deformation is downstream and downward.
(ii) Applied water load on the upstream foundation. Sherard (1973) comments that
the foundation has an important influence on the embankment deformation if it is
compressible or susceptible to collapse compression on wetting. Differential
foundation settlements can arise due to differential loading conditions from the
reservoir and in cases where the foundation differs under the embankment, such as
in the case of a cut-off trench to bedrock under the central core.
(iii) Decrease of effective stresses in the permeable earth and rockfill zones upstream
of the core. Deformations will be upward and upstream due to the unloading of
Chapter 7: Deformation behaviour of embankment dams Page 7.70

effective stresses, but are likely to be small due to the high tangent modulus on
unloading and reloading of gravels and rockfills compared to their modulus
during loading.

Figure 7.24: Effect of reservoir filling on a zoned embankment (Nobari and Duncan
1972b)

7.5.1 EFFECT OF WATER LOAD ON THE CORE

Assuming the central earthfill zone of the embankment behaves in undrained conditions
during first filling, the effect of reservoir filling will be for a net increase in total stress
acting on the upstream face of the core zone. The changes in total stress within the
central earthfill zone and downstream shoulder zones due to first filling will depend on:
• The magnitude of the change in total stress acting on the upstream face of the core
due to first filling, its location and the direction in which it acts as controlled by the
orientation of the upstream face of the earthfill zone.
• The existing stresses within the embankment at end of construction.

The magnitude of the total stress difference is greatest where the load acts on the
upstream slope of the embankment, as in the case of a homogeneous embankment or
embankments with facing elements designed as the main water barrier element (e.g.
concrete or asphalt faced rockfill dams). For typical slopes of homogeneous earthfill
embankments of greater than 2 - 3H to 1V, the applied direction of the water load is
dominantly vertical (Figure 7.25) and stress changes within the embankment will
Chapter 7: Deformation behaviour of embankment dams Page 7.71

mainly occur in the upstream slope. Resultant deformations of the crest or under the
embankment axis will therefore generally be small.

Figure 7.25: Water load acting on a homogeneous earthfill dam.

For zoned earthfill embankments with near vertical narrow cores and free draining
gravels or rockfill in the upstream shoulder, the direction of the applied water load is
near horizontal (Figure 7.24, Part 1). The magnitude of the difference in applied stress
on the upstream face from end of construction to reservoir full will be much less than
for water loads acting on the outer upstream face. It will be dependent on the existing
stress conditions within the embankment at end of construction, the hydrostatic stress
from the water and the reduction in effective stresses in the upstream rockfill on
reservoir filling.
The earlier finite difference analysis (Section 7.4.1.2, Figure 7.4) has been extended
to compare the lateral stress distributions at end of construction and end of first filling
for two cases:
• The central earthfill zone having the same compressibility properties as the rockfill
• A wet placed core of low undrained strength (Case 3 from the earlier analysis).

Given that a linear elastic perfectly plastic Mohr-Coulomb model has been used for
the embankment materials, the stress distributions will not necessarily be accurate at the
end of first filling but they will be indicative of the general trend.
Where the core has similar compressibility properties to the shoulder the lateral
stresses at end of construction are relatively low (Figure 7.26a) compared to the core of
low undrained strength (Figure 7.27a), as previously discussed in Section 7.4.1.2.
Assuming the core to behave undrained during reservoir filling, the change in total
lateral stress in the core and downstream shoulder on reservoir filling is shown to be
Chapter 7: Deformation behaviour of embankment dams Page 7.72

greater for the core with similar compressibility properties to the rockfill (Figure 7.26b).
For the low undrained strength core, lateral stress changes in the core and downstream
rockfill on reservoir filling are smaller in comparison (Figure 7.27b) because of the
already high lateral stresses present at the end of construction.
For both cases the lateral stress conditions in the core and downstream rockfill
increase so resultant lateral deformations will be in a downstream direction. The high
horizontal component of the applied stress vector from the water load acting on the
upstream face of the core, compared to its vertical component, results in greater changes
in lateral than vertical stress in the core and downstream shoulder. Therefore,
deformations will be dominantly horizontal. In reality settlements are observed on the
crest and downstream shoulder during first filling, most likely due to a combination of
time dependent creep and consolidation, and potentially due to collapse compression in
the downstream shoulder.
The magnitude of the deformation will depend on the changes in stress conditions,
the compressibility properties of the core and downstream rockfill and the embankment
geometry. For the downstream rockfill shoulder the existing lateral stress conditions are
shown to be important, but also important is the effective stress path on first filling,
which is for a net decrease in deviatoric stress and net increase in bulk stress.

Figure 7.26: Lateral stress distribution at (a) end of construction, and (b) reservoir full
condition, for central core earth and rockfill embankment with core of similar
compressibility properties to the rockfill.
Chapter 7: Deformation behaviour of embankment dams Page 7.73

Figure 7.27: Lateral stress distribution at (a) end of construction, and (b) reservoir full
condition, for a central core earth and rockfill embankment with wet placed core of low
undrained strength.

7.5.2 COLLAPSE COMPRESSION DURING FIRST FILLING

Collapse compression or collapse settlement on wetting of embankment materials is an


important factor in the deformation behaviour during first filling for a large number of
embankments. In most cases collapse compression is associated with wetting of the
upstream rockfill shoulder during first filling and numerous central core earth and
rockfill embankments exhibit this effect including El Infiernillo dam (Marsal and
Ramirez de Arellano 1967), Cougar dam (Pope 1967), Round Butte dam (Patrick 1967)
and Beliche dam (Naylor et al 1997) amongst many others. Collapse compression can
also occur in the downstream rockfill or gravel zones following heavy rainfall or due to
inundation of the downstream toe, and earthfills are also potentially susceptible.
Evidence of collapse compression during first filling at the main section includes
differential settlement across the crest with greater settlement of the upstream shoulder,
crest spreading and longitudinal crest cracking in some case studies.
The susceptibility of a rockfill to collapse compression is dependent on a number of
factors including; the rock substance unconfined compressive strength and its reduction
on saturation, moisture condition at placement (dry placed, water added or sluiced),
degree of compaction, particle size distribution (coarser rockfills are potentially more
susceptible) and applied stress levels prior to saturation. Case study and laboratory test
Chapter 7: Deformation behaviour of embankment dams Page 7.74

data (Nobari and Duncan 1972a; Marsal 1973; Alonso and Oldecop 2000) indicates the
potential for or magnitude of collapse compression is; greater for dry placed rockfills,
decreases with increasing compactive effort, and increases with increasing stress level
(refer Section 6.3.2 of Chapter 6). Dry dumped and poorly sluiced rockfills are
particularly susceptible to collapse compression as evidenced by the very large
deformations observed at Cogswell dam in 1933 following a very heavy rainstorm
(Baumann 1958) and the deformations observed at Strawberry and Dix River dams
(Howson 1939) on flooding during construction. Terzaghi (1960) attributed the
deformation to a reduction in compressive strength of the rockfill on saturation, mainly
in the outer surface of the particles, leading to failure at the highly stressed point
contacts in an angular rockfill mass.
Rock types that are generally more susceptible to collapse compression due to
greater strength reduction of the substance strength on saturation include sedimentary
rock types from sandstones down to mudrocks (mudrocks are potentially more
susceptible), limestones, and metamorphic rocks from sedimentary parent rocks.
Igneous rocks tend to be less susceptible.
Laboratory compression curves for a rockfill sample in dry and wet states (Nobari
and Duncan 1972a) are shown in Figure 7.28. Collapse compression on wetting occurs
for the dry rockfill when the stress state is above the normal compression line for wetted
or saturated rockfill, and the collapse strain is equivalent to the difference in strain
between the dry and wetted states at a given confining stress.
Earthfills are also susceptible to collapse compression but its occurrence is not as
frequently observed in modern embankment dams. This is mainly due to the high
compactive effort used and placement at moisture contents close to Standard optimum
moisture content.
If we consider the idealised results of a series of constant suction oedometer tests on
a compacted cohesive earthfill (Figure 7.29), collapse compression on wetting occurs
for a partially saturated soil when at a state, in void ratio – stress space, above that for
the saturated soil. At low stress levels heave will occur on saturation. A more typical
void ratio – stress relationship for a dry placed well-compacted cohesive earthfill
(placed within 1 to 2% dry of Standard optimum) is for a reduction in matric suction
with increasing stress (as shown) due to the compression of air voids in the soil
structure and increase in the degree of saturation. This is verified by the pore water
pressure response in piezometers during construction (Section 7.4.1.4). Therefore, this
Chapter 7: Deformation behaviour of embankment dams Page 7.75

reduction in matric suction reduces the susceptibility of the earthfill to collapse


compression on wetting compared to the constant suction condition.

Figure 7.28: Compression curves for dry and wet states and collapse compression from
the dry to wet state for Pyramid gravel in the laboratory oedometer test (Nobari and
Duncan 1972a).

Figure 7.29: Idealised effect of matric suction on collapse compression of earthfills


(Khalili 2002).

Another important aspect for well-compacted cohesive earthfills placed within 1 to 2%


dry of Standard optimum is that the air voids retained in the soil structure are not readily
removed on wetting to give complete saturation as they are for dry placed free draining
Chapter 7: Deformation behaviour of embankment dams Page 7.76

rockfills, sands or gravels. Therefore, on wetting the reduction in matric suction will
not be significant as a result of the retained air voids in the soil structure and collapse
compression, if any, will be limited. Sampling of the “moist placed” medium plasticity
compacted clay from below the phreatic surface at Belle Fourche dam (Sherard 1973)
more than 25 years after placement indicated the soil structure still retained an air voids
content of about 8%.
Factors that will affect the susceptibility of an earthfill to collapse compression are:
• Placement moisture content relative to Standard optimum. The drier soils are placed
the more susceptible they are to collapse compression on wetting.
• Compacted density. For a dry placed earthfill, the lower the density ratio of a given
soil type the more susceptible it is to collapse compression. The variation in density
within a compacted layer is also important. The lower portion of thick placed
layers, where the density is often lower, is more susceptible to collapse compression
than would be the upper, more heavily compacted part of the layer.
• Earthfill material type. For similarly dry placed earthfills, say at 3% dry of Standard
optimum, cohesionless and low cohesion earthfills appear, from the case study data,
to be more susceptible to collapse compression than cohesive earthfills.

A number of factors, other than those mentioned above, are likely to influence the
susceptibility to collapse compression including particle size distribution, soil plasticity,
degree of saturation and the ability of the partially saturated soil to retain air voids in the
soil structure on saturation.
Examples of collapse compression of earthfills are much more limited than for
rockfills in embankment dams. The database includes several suspected examples that
are discussed in Sections 7.9.2 and 7.11.1. The material types involved are mainly silty
to clayey sands and gravels that were placed well dry of Standard optimum moisture
content.
The effect of collapse compression of the upstream rockfill or gravel zone on the
embankment deformation behaviour during first filling is dependent on a number of
factors including the strain induced by collapse compression, the embankment zoning
and geometry, the strength and compressibility of the embankment materials and the
stress conditions at end of construction.
For earthfill cores of relatively low drained compressibility, typified by dry placed
well-compacted earthfills, collapse compression of the upstream rockfill on first filling
Chapter 7: Deformation behaviour of embankment dams Page 7.77

will result in greater settlement of the rockfill relative to the core (Figure 7.30). The
deformation at Cherry Valley dam (Lloyd et al 1958) is an example of this type of
deformation where the settlement of the upstream rockfill was more than four times that
of the central core and longitudinal cracks on the crest were observed at the upstream
core / rockfill interface. Shear stresses, acting as down drag on the upstream face of the
core, are developed at the upstream core / rockfill interface due to the differential
settlement (Squier 1970).

Figure 7.30: Cracking caused by post construction differential settlement between the
core and the dumped rockfill shoulders (Sherard et al 1963).

For zoned earth and rockfill embankments with wet placed cohesive earthfill cores the
deformation behaviour of the core during first filling is significantly different to that for
dry placed earthfills. On collapse compression of the upstream rockfill, the core is not
able to support the load that could be transferred to it as a result of differential
settlement (because of its low undrained strength). In effect the core is reliant on the
shoulders for support. Therefore, if the upstream shoulder settles due to collapse
compression on wetting then the core must deform with the upstream rockfill, largely as
plastic type deformations, and limited differential settlements occur at the upstream
interface.
The deformation behaviour of the wet placed clay core of El Infiernillo dam during
first filling (Marsal and Ramirez de Arellano 1967; Squier 1970; Nobari and Duncan
1972b) is an example of this type of deformation behaviour. Collapse compression of
the dry placed upstream rockfill occurred during first filling. The measured vertical
strains in the core were similar in magnitude and elevation to those in the upstream
rockfill, indicating that the core essentially deformed with the upstream shoulder. The
actual deformation behaviour at El Infiernillo dam is more complex than summarised
Chapter 7: Deformation behaviour of embankment dams Page 7.78

due to the influence of the upstream filters and effect of lateral spreading of the core.
The deformation behaviour at El Infiernillo dam is discussed further in Section 7.10 and
in Section 1.9 of Appendix G.
Two-dimensional finite difference analysis was undertaken to further assess the
effects of collapse compression of the upstream rockfill and the shear strength
properties of the core on the embankment deformation behaviour during first filling.
The modelled embankment (Figure 7.31) consisted a 100 m high central core earth and
rockfill dam with narrow core on a rigid bedrock foundation. Filters and zoning of the
rockfill shoulders were omitted for simplicity. Two different core types were modelled;
a very stiff dry placed core of low drained compressibility and high undrained strength,
and a wet placed core of low undrained shear strength that behaves undrained during
construction and on first filling. For each core type the upstream rockfill was modelled
with and without collapse compression. Similar properties were used for the
downstream rockfill in each case. The shear strength and compressibility properties of
the core and rockfill are given in Table 7.16 for the following four analysed cases:
• Case 8 – “dry” placed clay core of very stiff strength consistency and relatively low
drained compressibility with collapse compression in the upstream rockfill.
• Case 9 – “dry” placed clay core (as per Case 8), but without collapse compression in
the upstream rockfill.
• Case 10 – “wet” placed clay core of low undrained strength with collapse
compression in the upstream rockfill.
• Case 11 – “wet” placed clay core (as per Case 10), but without collapse compression
in the upstream rockfill.

Figure 7.31: Embankment model for finite difference analysis during first filling.
Chapter 7: Deformation behaviour of embankment dams Page 7.79

Embankment construction was modelled in ten stages each of ten metres thickness. A
Mohr-Coulomb linear-elastic perfectly plastic model was used for both the core and
rockfill during construction. The rockfill was modelled using the loading or
“construction” compressibility properties. At end of construction the deformations were
initialised to zero.
First filling was modelled assuming rapid filling of the reservoir such that the
permeable upstream rockfill was saturated and the core was effectively impermeable. A
load was applied to the upstream face of the core equivalent to the hydrostatic water
load and buoyant unit weights were used in the saturated portion of the upstream
rockfill. A reservoir full condition was at a height of 98 m above the foundation.
First filling was modelled differently depending on whether or not collapse
compression of the upstream rockfill was to be incorporated into the analysis. For the
cases where collapse compression was not included (Cases 9 and 11) first filling was
modelled in one stage. Changes to the material properties on first filling were:
• Upstream rockfill - buoyant unit weight and unloading compressibility properties of
the saturated portion.
• Downstream rockfill – increased shear modulus (the bulk modulus remained the
same) to reflect the effective stress path in the downstream shoulder, which shows
an increase in bulk stress and decrease in deviatoric stress on first filling.

For the cases incorporating collapse compression of the upstream rockfill (Cases 8
and 10) first filling was modelled in a series of 10 stages; the first 9 stages in 10 m
increments of rising reservoir level and the last stage as an 8 m increment to full supply
level. Collapse compression was modelled as per the method by Justo (1991), which
involved:
• For the newly saturated layer of each stage, collapse compression was modelled by:
− Reduction of the effective stresses within the newly saturated layer of the filling
stage. The reduction factor is equivalent to the compressibility ratio between the
“wetted” and “dry” states of the rockfill (Figure 7.32). In this case a ratio of 0.7,
or a 30% reduction, was used.
− Reduction of the Young’s Modulus of this layer to those for the “wetted”
rockfill in loading.
• For previously wetted layers, the unloading / reloading moduli were used for the
saturated upstream rockfill.
Chapter 7: Deformation behaviour of embankment dams Page 7.80

The downstream rockfill for Cases 8 and 10 was also modelled with an increased
shear modulus as per the modelling for Cases 9 and 11. In addition, the “dry” placed,
very stiff core for the model incorporating collapse compression of the upstream rockfill
(Case 8) was also modelled with an increased shear modulus. The ratio of the loading
to unloading moduli used was a factor of 10 for the rockfill and a factor of 5 for the
earthfill core.

Table 7.16: Properties of central core used in finite difference analyses


"Dry Placed" Core "Wet Placed" Core
Embankment Zone and Properties Case 8 Case 9 Case 10 Case 11
(collapse) (no collapse) (collapse) (no collapse)
Earthfill core:
- modulus, E (MPa)* 1 30 30 20 20
- Poisson’s ratio, ν 0.33 0.33 0.48 0.48
- cohesion, c' or cu (kPa) 5 5 50 50
- friction angle, φ' or φu (degrees) 32 32 2 2
- bulk unit weight, γ (kN/m3) 20 20 20 20
Rockfill:
- Poisson’s ratio, ν 0.25 0.25 0.25 0.25
- Cohesion, c' (kPa) 0 0 0 0
- Friction angle, φ' (degrees) 45 45 45 45
- Bulk unit weight, γbulk (kN/m3) 20 20 20 20
- Saturated unit weight, γsat (kN/m3) 22.3 22.3 22.3 22.3
- Young’s Modulus, E (MPa)
- loading (construction) 25 25 25 25
- “wet” modulus (on collapse) 17.5 na 17.5 na
- unloading / reloading for non-collapse. 250 250 250 250
- unloading / reloading post collapse. 175 175 175 175

1600
collapse compression on
wetting at constant stress
"dry" placed rockfill
Vertical Stress (kPa) .

1200

saturated rockfill
800

400
stress-strain relationship on
unloading and reloading

0
0.0% 2.0% 4.0% 6.0% 8.0%
Vertical strain (%)

Figure 7.32: Model of the stress-strain relationship of the upstream rockfill


incorporating collapse compression (in one-dimensional vertical compression).
Chapter 7: Deformation behaviour of embankment dams Page 7.81

The deformation results on first filling from the finite difference modelling are shown in
Figure 7.33 to Figure 7.36 for Cases 8 to 11 respectively. The settlement and
displacement of the embankment at the end of first filling are presented for all cases,
and for the reservoir at 40% of the embankment height for Case 10 only (wet placed
core modelling collapse compression of the upstream rockfill). It should be noted that
the deformations are not (and are not intended to be) an accurate representation of actual
deformation due to the use of simplistic models for the strength and compressibility of
the embankment materials. Importantly, though, the general trends of the deformation
behaviour are considered to be a reasonable representation of those in reality.
For the cases where collapse compression of the upstream rockfill was not modelled
(Cases 9 and 11) the results show:
• Settlements in the upstream rockfill and core are comparatively very much smaller
than those for the cases incorporating collapse settlement of the upstream rockfill.
• Upward vertical displacements are observed in the mid to lower upstream shoulder
due to decreases in the effective vertical stress on first filling. These are obscured in
the collapse compression analyses (Cases 8 and 10) due to the collapse compression.
• Displacements are downstream in the core and shoulders.

For the cases where collapse compression of the upstream rockfill was modelled
(Cases 8 and 10) the results show that whilst the settlements in the upstream shoulder
are similar for the different core types, the settlements within the core show markedly
different patterns of behaviour. For the “dry” placed core model (Case 8), the
settlement trend shows (Figure 7.33a) a large differential settlement at the upstream
interface of the core. This pattern of settlement is similar to that observed from the case
studies, except that the differential settlement in the crest region is often more
pronounced in the case studies. This is partly a result of the relatively large shear type
deformation that has developed within the core in the model. By adopting higher shear
strength parameters for the core, greater differential settlement occur at the upstream
core / rockfill interface and smaller settlements within the core.
In contrast, the numerical model for the “wet” core of low undrained shear strength
shows virtually no differential settlement at the upstream core interface (Figure 7.35a
and c), indicating the core is deforming with the upstream rockfill shoulder as observed
in case studies with wet placed clay cores such as El Infiernillo and Canales dams
(Giron 1997, refer Section 2.1 of Appendix G). Instead, the modelling shows a
Chapter 7: Deformation behaviour of embankment dams Page 7.82

concentration of vertical settlement in the lower part of the core and a large differential
settlement at the downstream interface of the core. The concentration of deformation at
the downstream interface of the core is clearly demonstrated at Canales dam.

Figure 7.33: Case 8 – “dry” placed, very stiff core modelling collapse compression in
the upstream rockfill; numerical results of (a) vertical and (b) lateral displacement on
first filling.

Figure 7.34: Case 9 – “dry” placed, very stiff core without collapse compression;
numerical results of (a) vertical and (b) lateral displacement on first filling.
Chapter 7: Deformation behaviour of embankment dams Page 7.83

Figure 7.35: Case 10 – “wet” placed clay core of low undrained strength modelling
collapse compression in the upstream rockfill. Numerical results of vertical and lateral
deformation on first filling; (a) and (b) for reservoir at 40% of embankment height, (c)
and (d) for reservoir at 98% of embankment height.
Chapter 7: Deformation behaviour of embankment dams Page 7.84

Figure 7.36: Case 11 – “wet” placed clay core without collapse compression; numerical
results of (a) vertical and (b) lateral displacement on first filling.

The lateral displacement of the core on first filling shows a similar “bow” shaped
pattern for all four cases, with the region of maximum downstream displacement on the
dam axis occurring in the mid region of the core. The displacement for the cases
incorporating collapse compression show a tendency for regions of the core and
upstream shoulder to displace upstream. For the “wet” placed core (Case 10) the lateral
deformation of the core was largely upstream for the reservoir at 40% of the
embankment height (Figure 7.35b) and then largely downstream at the end of first
filling (Figure 7.35d), still retaining its bow shape. This upstream then downstream
deformation pattern of the crest is not uncommon in zoned embankments with narrow,
near vertical central cohesive cores placed wet of Standard optimum; it was observed at
El Infiernillo, Thomson, Gepatsch and Nurek dams.

7.5.3 LATERAL SURFACE DEFORMATION NORMAL TO THE DAM AXIS DURING FIRST
FILLING.

As discussed in Section 7.5.1, a significant portion of the post construction lateral


displacement occurs during first filling, particularly for embankments with thin to
medium width central cores and permeable upstream fill (mainly central core earth and
rockfill dams). The lateral surface displacements during first filling for the case studies
Chapter 7: Deformation behaviour of embankment dams Page 7.85

of earthfill and zoned earth-rockfill embankments are presented in the following figures
and tables.
For analysis of the post construction surface deformation, the embankment has been
divided into three regions (Figure 7.37); the mid to downstream region of the crest, the
mid to upper region of the downstream slope and the upper upstream slope to upstream
crest region. The regional division was primarily established to separate the
deformation of the core of zoned embankments from that of the upstream and
downstream shoulders due to the sometimes large difference in observed deformation
behaviour, particularly where poorly compacted rockfill was used in the shoulders. It is
recognised that because the deformation of the shoulders affects that of the core, any
assessment of the embankment deformation requires consideration of the embankment
as a whole, but given the large number of case studies within the database it is
impractical to do so for all dams. In the discussion on individual embankments in
Sections 7.10 and 7.11, and in Appendix G the deformation of the embankment as a
whole is discussed for selected case studies.

Figure 7.37: Regional division of the embankment for analysis of post construction
surface deformation.

Representative surface monitoring points (SMP) for an embankment were generally


taken from near to the maximum section, unless otherwise identified. Other selection
criteria for SMPs are:
• For the downstream slope SMPs were selected from the slope region in the range
0.6 to 0.8 times the embankment height, which is often where the greatest
Chapter 7: Deformation behaviour of embankment dams Page 7.86

displacements were measured. Embankments were not excluded if the only SMPs
on the downstream slope were outside this region.
• For zoned embankments with thin to medium width cores SMPs representing the
upper upstream shoulder to upstream crest region were preferentially selected from
the upstream edge of the crest provided the SMP was located over the upstream
shoulder zone.
• For embankments with very broad earthfill zones (including earthfill embankments)
the crest region was considered as a whole; i.e. from the upstream edge to the
downstream edge. For these embankments SMPs from only the upper upstream
slope region were included with the analysis of the upper upstream slope to
upstream crest region.

Figure 7.38 to Figure 7.40 present the case study results for the lateral surface
displacements during first filling:
• Figure 7.38 for the mid to downstream region of the crest. The data has been sorted
based on the width of the central core, and the material type and compaction of the
downstream shoulder.
• Figure 7.39 for the mid to upper region of the downstream slope, sorted based on the
material type and compaction of the downstream shoulder.
• Figure 7.40 for the upper upstream slope and upstream crest region, sorted based on
material type and compaction of the upstream shoulder.

The terms “well-compacted”, “reasonably to well-compacted”, “reasonably


compacted” and “poorly compacted” used to describe the compaction of the rockfill in
the embankment shoulders are defined in Section 1.3.3. The term “gravels” refers to
free draining coarse grained gravelly and bouldery soils with very low fines content
(less than about 5 to 8% passing 75 micron). The term “earthfill” covers a broad range
of material types from clays to silty gravels to weak rockfills that breakdown
significantly on compaction.
Zero time for measurement of displaceme nt, t0, is established at the end of
embankment construction. In most cases the displacement for each case study captures
a large portion of the period of first filling, however, for several cases where the
embankment impounded water to relatively high levels during construction some of the
displacement due to first filling would have been missed (e.g. Vatnedalsvatn dam). No
Chapter 7: Deformation behaviour of embankment dams Page 7.87

differentiation with respect to time of first filling is taken into consideration in the
figures. In some cases the reservoir was filled within a matter of days to weeks whilst
for others first filling took many years, more than 10 years for several cases.
Highlighted in the figures are those embankments where the rockfill was known to
have been placed without the addition of water. These cases are highlighted because of
the susceptibility of dry placed rockfill, particularly if not well compacted or if
weathered, to collapse compression on wetting. It is recognised that for several of these
highlighted embankments the rockfill was likely to have been wetted from rainfall in
regions where the climate is very wet, but they have been included rather than to make
this judgement.
The figures present information on the general trend and magnitude of the lateral
displacement on first filling, and highlight case studies with “abnormally” large
deformation or with trends different to the broader group. A more detailed analysis of
the overall post construction lateral deformation behaviour is presented in Section 7.6
and the discussion on individual embankments in Sections 7.10 and 7.11.

(a) Crest displacement on first filling


For the displacement of the crest (Figure 7.38) the material forming the downstream
shoulder and the width of the core region do have some influence (refer below) but
generally speaking it is difficult to establish bounds of deformation other than a general
range. Reasons for this are considered to include factors such as the time of first filling,
the stress conditions within the downstream shoulder prior to first filling (refer Section
7.5.1), the effect of collapse settlement of the upstream rockfill and resultant core
deformation (refer Section 7.5.2), positioning of the SMP on the crest, valley shape, and
the curvature of the embankment axis amongst other factors. Attempts to further
differentiate the data for one or more of these aspects often resulted in too great a
dilution of the database to draw any significant conclusions.
Table 7.17 presents a summary of the lateral crest displacement on first filling. The
displacement range for most case studies is from 50 mm upstream to 300 mm
downstream (Figure 7.38). For most groups in Table 7.17 the range is less than about
100 to 200 mm, or less than 0.1 to 0.2% of the embankment height. Greater
displacements are observed for dams with reasonably to poorly compacted rockfill in
the downstream shoulder or embankments with cores of silty sands to silty gravels and
of thin to medium width.
Chapter 7: Deformation behaviour of embankment dams Page 7.88

1200
1% Svartevann 0.5%

900
(downstream is positive) .
Displacement (mm) .

0.25%
600

300

indicates rockfill placed without


Mita Hills addition of water
Rector Creek
-300
0 50 100 150 200 250 300
Dam Height (m)
Thin to medium cores, well-compacted rockfill Thick cores, reasonably to well compacted rockfill
Thin to medium cores, reasonably to well compacted rockfill Thick cores, poor compaction of rockfill
Thin to medium cores, reasonable compaction of rockfill Thick cores, dominantly gravels shoulders
Thin to medium cores, poor compaction of rockfill Thick cores, dominantly earthfill shoulders
Thin to medium cores, dominantly gravels shoulders Very broad cores
Very broad cores, possible foundation influence

600
0.5% 0.2%
La Grande 2
500
Djatilihur
(downstream is positive) .

Round Butte
Displacement (mm) .

400 indicates rockfill placed without


Beliche Cougar addition of water

300 Glenbawn
0.1%
Horsetooth
200

100

0
El Infiernillo
San Luis (slide area) South Holston -0.05%
-100
0 50 100 150 200 250 300
Dam Height (m)
Figure 7.38: Lateral displacement of the crest (centre or downstream region) on first
filling versus embankment height (displacement is after the end of embankment
construction).
Chapter 7: Deformation behaviour of embankment dams Page 7.89

Table 7.17: Lateral displacement of the crest (centre to downstream edge) on first filling.
Core Downstream Shoulder Core *2 No. Displacement Range
Comments
Width Material Rating * 1
Classification Cases (mm) % of H *3
CL/CH/GC/SC, wet 13 29 to 180 0.0 to 0.12
Well-comp CL/SC/GC, dry 5 5 to 80 0.0 to 0.12
SM/GM – dry and wet 3 177 to 394 0.10 to 0.30 Naramata (177 mm), Cougar (375 mm), Round
Butte (394 mm)
CL/CH/SC/GC 5 6 to 78 0.0 to 0.11
Thin to Rockfill Reas to well
SM/GM 5 6 to 535 0.03 to 0.36 LG2 (535 mm), Frauenau (250 mm)
medium
Reas All types 4 -6 to 1120 -0.01 to 0.17 Svartevann = 1120 mm (SM core)
(most < 200)
No correlation to core soil type, Beliche =
Poor All types 7 -39 to 480 0.09 to 0.63
347 mm (0.63%), Djatiluhur = 480 mm (0.46%)
Gravels - All types 3 18 to 107 0.04 to 0.07
-64 to 285
All cases with thick cores 19 -0.02 to 0.20
(most –19 to 222)
Reas to well All types 1 0 - Maroon dam
Rockfill Poor All types 9 -64 to 285 -0.02 to 0.19 Sth Holston = -64 mm (-0.07%),
Thick
(most –15 to 165) Glenbawn = 285 mm (0.37%)
Gravels - All types 4 -13 to 44 -0.01 to 0.04
Earthfill - All types 5 -19 to 222 -0.02 to 0.19 Navajo = 222 mm, rest < 45 mm
(most –19 to 45) (-0.02 to 0.09)
-236 to 229 Rector Creek = -236mm, Mita Hills = -146mm,
Very Broad All types 15 -0.02 to 0.14 San Luis (slide area) = -79 mm, Horsetooth =
(most –58 to 94)
229 mm
1
* compaction rating of rockfill; well-comp = “well-compacted”, reas to well = “reasonably to well compacted”, reas = “reasonable compaction”, poor = “poorly compacted”.
*2 Soil classification to Australian Standard AS 1726-1993, “wet” = placed at or on the wet side of OMC, “dry” = placed on the dry side of OMC (OMC = Standard optimum
moisture content).
*3 range of displacement as a percentage of dam height (H) excludes possible outliers (Svartevann, South Holston, Glenbawn, Rector Creek, Mita Hills, San Luis (slide area)
and Horsetooth dams).
Chapter 7: Deformation behaviour of embankment dams Page 7.90

Isolated cases of large displacement also occur within the other groupings. For the
embankments with cores of very broad width, the relatively large upstream
displacement for Rector Creek and Mita Hills is possibly due to collapse compression of
dry placed dominantly sandy earthfills, for the slide area at San Luis dam due to the
upstream trending hill slope on which this section of the embankment was constructed,
and for Horsetooth dam the foundation is thought to have had a significant influence on
the displacement. These cases are discussed further in Section 7.11. South Holston
would also appear to be an outlier in its sub-grouping with an upstream displacement of
64 mm (-64 mm) on first filling. At Glenbawn dam the crest SMP is located at the
downstream edge over the poorly compacted rockfill zone, which would account for the
large displacement measured.
Further details on the embankments comprising the 13 largest downstream
displacements on first filling (excluding the very broad earthfill case studies) are
summarised in Table 7.18. All have a lateral displacement equal to or greater than 200
mm downstream on first filling. These case studies highlight several factors associated
with high lateral displacements:
• The case studies are predominantly of central core earth and rockfill dams with thin
cores.
• The predominant core type is silty sands to silty gravels, particularly for the larger
deformations (4 of the top 5).
• For most cases (8 of 13) the downstream rockfill is of poor to reasonable
compaction.
• In 7 cases the downstream rockfill was known to have been placed without the
addition of water. Of these, 2 of the 3 well or reasonably to well compacted cases
were placed without water, all the reasonable compaction cases were placed without
water and 3 of the 4 poorly compacted cases were placed without the addition of
water.

The data indicates that potentially important factors leading to relatively large
downstream displacement on first filling are:
• Collapse settlement of dry placed or poorly compacted rockfills in the downstream
shoulder, possibly as a result of post construction wetting from heavy rainfall.
• Relatively high increases in stress level in the downstream rockfill from the applied
water load. The earlier finite difference analysis for zoned embankments (Section
Chapter 7: Deformation behaviour of embankment dams Page 7.91

7.4.1.2, Figure 7.4 and Section 7.5.1, Figure 7.26) showed that where the
compressibility properties of the core are similar to that of the shoulders the lateral
stresses are relatively low at end of construction, and that on first filling the
magnitude of increase in lateral stress in the core and downstream shoulder is
relatively high. Following on from this, it is reasonable to consider that potentially
larger downstream displacements occur for earth and rockfill embankments with
thin to medium width earthfill cores of well-compacted silty sands to silty gravels.
Although not conclusive, it may be a factor leading to large displacement on first
filling.

(b) Displacement of the downstream shoulder on first filling


A similar pattern of behaviour as the crest is observed for the lateral displacement of the
mid to upper region of the downstream slope on first filling (Figure 7.39). The data
shows that:
• For embankments with very broad earthfill zones or earthfills in the downstream
shoulder, displacements are typically less than 100 mm downstream (or less than
0.1% to 0.15% of the embankment height). The much greater displacement at
Horsetooth dam is due to the influence of the compressible foundation.
• For embankments with gravels in the downstream shoulder, displacements are
typically less than 100 to 150 mm (or less than 0.11% of the embankment height).
The higher displacement at Trinity dam (175 mm, 0.11%) may reflect the use of
dozer track rolling and the limited amount of water used in compaction of the
gravels.
• For embankments where wetted rockfills are used in the downstream shoulder,
displacements are typically less than 100 to 200 mm (or less than 0.12 to 0.18% of
the embankment height). The displacements at Ataturk and Beliche dams are clear
outliers to the general trend, and Blowering and Windemere are slightly above the
general trend. At Beliche, Blowering and Ataturk dams the higher displacements
are possibly due to the use of weathered materials and/or rockfills susceptible to
strength loss on wetting. At Windemere placement in 2 m layers and possibly
without the addition of water in the outer downstream slope may be a factor.
Chapter 7: Deformation behaviour of embankment dams Page 7.92

Table 7.18: Thirteen cases with largest downstream crest displacement on first filling
Downstream Core *3 Downstream Shoulder Lateral Time to First
Height % of First
Dam Name Slope *2 Moisture Rating Displacement Filling
(m) *1 Width Classn Material Placement *5 6 Filling *7
(H to 1V) Content *4 (mm) (years) *
Svartevann 129 1.43H Thin (u) SM/GM 0.4% wet Rockfill – granite & reas 2m lifts, no water, 8p 1120 2.23 30%
gneiss 13t SDVR
La Grande 2 150 1.6H Thin (u) SM (1% dry to 2% Rockfill – granite reas to 0.9 to 1.8m lifts, no 535 1.2 100%
wet), wet ? well water, 10t SDVR
Djatiluhur 105 1.3H (upper) Thin CH 2.2% wet Rockfill – andesite poor 1 to 2 m lifts, dumped, 480 2.25 30%
watered
Round Butte 134 1.7H Thin SM dry ? Rockfill – basalt well 0.6m lifts, no water, 4p 395 0.6 18%
9t SDVR
Cougar 136 1.6H Thin (u) GM 1.0% wet Rockfill – basalt & well 0.6m lifts, 4p 10t 375 0.67 most
andesite SDVR
Beliche 55 1.95H Thin GC (± OMC), Rockfill – schists & poor 1m lifts, light 345 4.1 50%
wet ? greywacke compaction
Glenbawn 76.5 2.5H Thick CL 2% dry Rockfill – limestone poor 1.8m lifts, dumped, no 285 4.7 100%
water
Mud Mountain 124 1.75H Medium SM/GM dry ? Rockfill – andesite poor 12m lifts, dumped & 275 >17 most
& tuff sluiced
Frauenau 80 1.6H Thin SM dry ? Rockfill - No information 250 - most
Matahina 80 2.3H Medium CL 1 to 1.5% dry Rockfill – poor 1.0m lifts, no water, 230 0.24 100%
ignimbrite dumped
Navajo 120 2.5H Thick CL/ML 0.8% dry Earthfill - 220 11 most
Gepatsch 153 1.5H Thin GM/GC 0.5 to 2% wet Rockfill - gneiss reas 2.0m lifts, no water, 4p 200 1.92 30%
9t SDVR
Vatnedalsvatn 121 1.5H Thin SM 0.5% wet Rockfill – quartz, reas Reas (1.5m lifts, no 200 0.2 10%
granite & gneiss water, 6p 11t SDVR)
Notes to Table 7.18 on next page
Chapter 7: Deformation behaviour of embankment dams Page 7.93

Notes to Table 7.18:


*1 dam height at SMP (surface measuring point)
*2 downstream slope angle, horizontal to vertical. Values given are to 1V.
*3 (u) indicates core oriented slightly upstream. Placement moisture content relative to Standard
optimum. Values in brackets are specification, ? indicates not sure but likely from pore water
pressure response.
*4 compaction rating of rockfill; well-comp = “well-compacted”, reas to well = “reasonably to well
compacted”, reas = “reasonable compaction”, poor = “poorly compacted”.
5
* shoulder placement. 8p 13t SDVR = 8 passes of a 13 tonne smooth drum vibrating roller.
*6 time to first filling after the end of embankment construction.
*7 percent of first filling is a reference to the reservoir level prior to the end of construction. 30% means
the reservoir was filled to within 0.3H (H = dam height) of the reservoir full condition. Therefore low
percentages indicate the reservoir was or had been close to full supply level prior to the end of
construction and high percentages or “most” indicates little water was impounded during construction.

• For embankments where dry placed rockfills are used in the downstream shoulder,
displacements are typically in the order of 150 to 900 mm, or in the range 0.10 to
0.85% of the embankment height. Collapse compression on wetting due to
rainwater infiltration is possibly a significant factor in the large displacements
observed. Notably, greater displacements are observed for the dry placed and
reasonably to poorly compacted rockfills (greater than 0.2 to 0.25%), which will be
more susceptible to collapse compression on wetting. For the dry placed, well and
reasonably to well compacted rockfills displacements are generally less than 0.25 to
0.30% of the embankment height.

Case studies with large downstream displacement of the crest also had large
downstream displacement of the downstream slope. The exception is Navajo dam (crest
displacement = 222 mm or 0.19%, downstream slope displacement = 90 mm or 0.07%)
where earthfill was placed in the downstream shoulder. Other notable cases with large
downstream displacement include:
• Eppalock dam, a 47 m high central core earth and rockfill dam with a medium
plasticity sandy clay core of thick width placed on the dry side of Standard optimum
(refer Section 1.10 of Appendix G). The shoulders were of dry placed and poorly
compacted rockfill (placed in 2 to 4 m thick layers and spread by tractor) derived
from basalt, but was weathered and contained a high fraction of finer sized rockfill.
The displacement of the downstream slope on first filling (290 mm or 0.63% of the
dam height) was much greater than for the crest (90 mm or 0.2% of the dam height).
Chapter 7: Deformation behaviour of embankment dams Page 7.94

• Eildon dam, an 80 m high central core earth and rockfill dam with a well-compacted
medium plasticity sandy to gravelly clay core of thick width placed on the dry side
of Standard optimum (refer Section 1.8 of Appendix G). The shoulders were of dry
placed and poorly compacted rockfill (placed in 2 m thick layers with no formal
compaction) sourced predominantly from quartzitic sandstone. The outer
downstream random rockfill shoulder consisted of “unsuitable” rock that was poorly
graded and contained a high fraction of finer sized rockfill. The displacement of the
downstream slope on first filling was very large at 660 mm (or 0.83% of the dam
height). The crest displacement was not monitored during first filling.
• El Infiernillo, Dartmouth, Thomson and Blowering dams also show much greater
displacement of the downstream slope than the crest. All comprise wet placed
clayey sand to sandy clay cores of thin to medium width. For these embankments
the core is likely to be strongly reliant on the shoulders for support due to its likely
low undrained strength and thin to medium width. Therefore, plastic deformation is
likely to be a significant factor in the deformation behaviour of the core during first
filling of these embankments affected by collapse settlement of the upstream rockfill
shoulder on wetting, as described for El Infiernillo dam (after Nobari and Duncan
1972b) in Section 7.5.2. Thomson and Dartmouth dams show an upstream then
downstream displacement during first filling similar to that at El Infiernillo dam.

At Eildon and Eppalock dams, embankments with thick cores of high undrained
strength and relatively low drained compressibility, the displacement records suggest
that the displacement of the dry placed, poorly compacted downstream rockfill shoulder
is not likely to be greatly influenced by first filling. The large displacement of the
downstream slope for these cases is more likely due to the large differential settlement
between the well-compacted core and downstream rockfill, and the influence of the
downstream slope of the core on the vector of deformation. Collapse compression in
the downstream rockfill due to wetting from rainfall would contribute to the large
deformations of the downstream slope observed during first filling. Similar reasoning
would also explain the differential displacement at Copeton dam (crest displacement =
75 mm, downstream slope = 360 mm) where the medium width clayey sand core was
dry placed and no water used during placement of the rockfill.
The deformation behaviour of several of these embankments, including Ataturk
dam, are discussed in more detail in Section 7.10 and in Appendix G.
Chapter 7: Deformation behaviour of embankment dams Page 7.95

1500
Well-compacted rockfill 1%
Reasonably to well compacted rockfill Ataturk
Reasonably compacted rockfill
1200
Poorly compacted rockfill
(downstream is positive) .

gravels
Displacement (mm) .

Earthfill 0.5%
900 earthfill, potential foundation influence
Svartevann

Eildon
600
0.25%

300

0
indicates rockfill placed without
addition of water
-300
0 25 50 75 100 125 150 175 200
Dam Height (m)

600
0.50%

indicates rockfill placed without Vatnedalsvatn


500 addition of water
(downstream is positive) .

Thomson
Displacement (mm) .

400
Copeton 0.20%
Gepatsch
Eppalock
300
Glenbawn Blowering
Beliche Round Butte Dartmouth
Trinity
200
Horsetooth 0.10%
Windemere

100
El Infiernillo

-0.05%
-100
0 25 50 75 100 125 150 175 200
Dam Height (m)
Figure 7.39: Lateral displacement of the downstream slope (mid to upper region) on
first filling versus embankment height (displacement is after the end of embankment
construction).

(c) Displacement of the upstream shoulder on first filling


The displacement of the upper upstream slope or upstream crest region during first
filling (Figure 7.40) highlights Svartevann and Gepatsch dams as outliers to the general
displacement trend. For most cases studies the displacement is in the range 200 to 300
mm upstream (-200 to -300 mm) to 300 mm downstream.

(d) General trends of displacement on first filling


Chapter 7: Deformation behaviour of embankment dams Page 7.96

The displacement on first filling of the crest and downstream slope were also plotted
against the downstream slope (Figure 7.41 and Figure 7.42). The plots indicate a
general trend of decreasing magnitude of displacement with flattening of the
downstream slope, with some outliers of cases with dry placed rockfill in the
downstream shoulder. However, this trend is only evident at the upper bound of lateral
displacement as a large number of case studies with relatively steep downstream slope
angles show limited deformation on first filling.
The general trends of lateral displacement on first filling are:
• For the crest (centre to downstream region), displacements typically range from 50
mm upstream (-50 mm) to 200 mm downstream, or from -0.02% to 0.20% of the
embankment height.
• For the downstream slope (mid to upper region), displacements are typically
downstream in the range from 0 up to 200 to 250 mm (or less than 0.2% of the
embankment height).
• In the upper upstream slope and upstream crest region, displacements typically
range from 200 mm upstream (-200 mm) to 200 mm downstream. The mean is
located closer to zero in this region than for the mid to downstream region of the
crest.

Much larger downstream displacements on first filling are observed:


• In the crest region for zoned earth and rockfill embankments with thin cores and dry
placed or poorly compacted rockfill in the downstream shoulder. In these cases
displacements can reach up to 0.3 to 0.65% of the embankment height (up to 600
mm), and in one case 1100 mm or (0.87%) was measured.
• In the mid to upper region of the downstream slope for zoned embankments with
dry placed or poorly compacted rockfill in the downstream shoulder. For these
cases displacements range from 150 mm to 900 mm (or from 0.10 to 0.85% of the
embankment height).

Overall the plots provide useful information on likely bounds of lateral displacement
on first filling. They are also useful for identification of embankments that display
excessively large displacements on first filling and therefore potentially “abnormal”
deformation behaviour. Several of the outliers are discussed further in Sections 7.9 to
7.11, and in Appendix G.
Chapter 7: Deformation behaviour of embankment dams Page 7.97

1200
Svartevann Well-compacted rockfill
Reasonably to well compacted rockfill
(downstream is positive) . 900 0.50% Reasonably compacted rockfill
Poorly compacted rockfill
Displacement (mm) .

gravels
600 Earthfill
0.25% earthfill, potential foundation influence
rockfill placed without addition of
300
water

-300
Glenbawn

-0.50% Gepatsch -0.25%


-600
0 50 100 150 200 250 300
Dam Height (m)

300 0.25%
Cougar 0.10%
Eildon Vatnedalsvatn
200
Round Butte
Horsetooth
(downstream is positive) .
Displacement (mm) .

100

0
Well-compacted rockfill
Reasonably to well compacted rockfill
-100 Reasonably compacted rockfill
Poorly compacted rockfill
San Luis -0.10% gravels
-200
(slide area) Dixon Canyon
La Agnostura Earthfill
earthfill, potential foundation influence
-300
rockfill placed without addition of
Glenbawn
water
-0.25%
-400
0 50 100 150 200 250 300
Dam Height (m)

Figure 7.40: Lateral displacement at the upper upstream slope of upstream crest region
on first filling versus embankment height (displacement after the end of embankment
construction).
Chapter 7: Deformation behaviour of embankment dams Page 7.98

1200
Svartevann indicates rockfill placed without
addition of water

900
(downstream is positive) .
Displacement (mm) .

600
La Grande 2
Djatiluhur

300

San Luis (slide area)


Mita Hills
Rector Creek
-300
1 1.5 2 2.5 3 3.5
Downstream slope (horizontal to vertical)
Thin to medium cores, well-compacted rockfill Thick cores, reasonably to well compacted rockfill
Thin to medium cores, reasonably to well compacted rockfill Thick cores, poor compaction of rockfill
Thin to medium cores, reasonable compaction of rockfill Thick cores, dominantly gravels shoulders
Thin to medium cores, poor compaction of rockfill Thick cores, dominantly earthfill shoulders
Thin to medium cores, dominantly gravels shoulders Very broad cores
Very broad cores, possible foundation influence

Figure 7.41: Lateral displacement of the crest (centre to downstream region) on first
filling versus downstream slope.

1500
Well-compacted rockfill
Ataturk Reasonably to well compacted rockfill
1200 Reasonably compacted rockfill
Poorly compacted rockfill
(downstream is positive) .

gravels
Displacement (mm) .

900 Earthfill
Svartevann earthfill, potential foundation influence

Eildon indicates rockfill placed without


600 addition of water

300

-300
1.0 1.5 2.0 2.5 3.0 3.5
Downstream slope (horizontal to vertical)

Figure 7.42: Lateral displacement of the downstream slope (mid to upper region) on
first filling versus downstream slope.
Chapter 7: Deformation behaviour of embankment dams Page 7.99

7.6 GENERAL P OST CONSTRUCTION D EFORMATION BEHAVIOUR OF


EARTHFILL AND ZONED EARTH AND EARTH -ROCKFILL
EMBANKMENTS
The post construction deformation data collected and presented for earthfill and earth-
rockfill embankments includes:
• Surface monitoring point (SMP) data on the crest and slopes of the embankment.
Both the vertical and horizontal deformation (in the plane normal to the dam axis) is
presented.
• Internal vertical deformation data, mainly from IVM gauges, in the central core
region of the embankment.

The data presented is usually for the maximum section of the embankment.
Zero time, t0, is established at the end of embankment construction to provide a
consistent basis point for comparison. The unit of time used for all data is years.
Where possible the deformation records used are the actual records. This has been
possible for a large number of the Australian dams and most of the United States
Bureau of Reclamation (USBR) dams. For a number of case studies however the data
has been digitised from published plots in the literature, and for these cases the
deformation as plotted is only a representation of the actual measured data as opposed to
the actual data records.
The data is presented mostly in graphical format with data or case studies sorted into
what are considered appropriate groupings. The number of plots presented is numerous,
particularly for the deformation versus time plots, due to the large number of cases and
number of groupings used. Total deformation plots incorporating a large number of
case studies are also presented at selected time intervals and these visually provide a
broad representation of the database.
The data for analysis of the post construction surface deformation behaviour is
divided into the three regions of the embankment as previously identified in Section
7.5.3 (Figure 7.37). They are; the mid to downstream region of the crest, the mid to
upper region of the downstream slope, and the upper upstream slope to upstream crest
region.
Chapter 7: Deformation behaviour of embankment dams Page 7.100

7.6.1 POST CONSTRUCTION INTERNAL VERTICAL DEFORMATION OF THE CORE

Analysis of the post construction internal vertical deformation of the core was
undertaken for those embankments for which detailed results of the internal deformation
monitoring were made available. This was generally limited to a number of Australian
and USBR dams, but includes several dams published in the literature.
The post construction IVM records were analysed by plotting the cumulative post
construction settlement or vertical strain between the individual cross-arms.
Assessment of the behaviour was based on the shape of the cumulative settlement or
vertical strain profile at a specific time interval after construction and over time.
Typical plots of cumulative settlement are presented in Figure 7.43 and Figure 7.44
for Talbingo (IVM ES1) and Copeton (IVM B) dams respectively. They are considered
representative of “normal” type deformation behaviour.
For Talbingo dam the much higher post construction vertical strains in the lower 40
to 50 m of the 160 m high dam (average of 0.7% at 24 years post construction) are
considered to be related to consolidation due to dissipation of high pore water pressures.
On first filling, pore water pressures increased by some 200 to 400 kPa above the
already high construction pore water pressures. Over the next 20 plus years the pore
water pressures slowly dissipated, reducing by as much as 600 to 1100 kPa below the
levels at end of first filling. In the mid region of the core the change in pore water
pressure due to dissipation was much less than in the lower section and is reflected in
the lower vertical strains (average of 0.3% at 24 years post construction). The localised
zone of high settlement between 55 and 58 m below crest level is potentially indicative
of “abnormal” behaviour. From the data it is evident that the differential settlement
between these IVM gauges occurred in the period between 0 and 0.15 years post
construction and prior to first filling, and since then has increased only marginally. The
reason for the behaviour is not known, but given its timing in relation to end of
construction and first filling it is not likely to be “abnormal”.
The behaviour of IVM B at the 110 m high Copeton dam is similar to that at
Talbingo dam. Higher vertical strains (average of 1.1% at 26 years post construction)
were observed in the lower 10 m of the core where pore water pressure dissipation post
first filling was in the order of 450 kPa. In the mid to upper regions of the core pore
water pressures were negligible at end of construction and slowly rose post first filling.
Vertical strains in this region of the core were on average 0.43% at 26 years post
construction, much less than in the lower 10 m of the core.
Chapter 7: Deformation behaviour of embankment dams Page 7.101

20

40
Depth below crest (m) .

60

80

100

ES1 - 1.2 years post EOC


120
ES1 - 6.6 years post EOC
ES1 - 9 years post EOC
140 ES1 - 24 years post EOC

Base of ES1 at 153 m depth EOC = end of embankment construction


160
0 100 200 300 400 500 600 700
Settlement (mm)

Figure 7.43: Post construction internal settlement of the core at Talbingo dam (IVM
ES1 at the main section).

27 June 1973
10 (0 yrs post EOC)

20

30
Depth below crest (m) .

40

50

60

70

80 July 1975 - 2.1 years post EOC


Nov 1978 - 5.4 years post EOC
June 1986 - 13 years post EOC
90
Mar 1999 - 25.7 years post EOC
EOC = end of embankment construction
100

110
0 200 400 600 800
Settlement (mm)

Figure 7.44: Post construction internal settlement of the core at Copeton dam (IVM B, 9
m downstream of dam axis at main section).
Chapter 7: Deformation behaviour of embankment dams Page 7.102

In contrast, the behaviour of IVM A at the 55 m high Bellfield dam (Figure 7.45),
located in the thick central clay core, is different. The IVM records from end of
construction to February 1987 indicate two potential zone of high localised vertical
strain. The records from February 1987 to November 1997 (Figure 7.45b) show a
localised zone of higher vertical strain developing at 28 to 30 m below crest level,
inconsistent with the localised zones of high strain prior to 1987. The narrow rockfill
shoulders at Bellfield were dry placed and poorly compacted, and based on the post
construction behaviour of Eppalock dam (of similar design and construction to
Bellfield), the “abnormal” behaviour may represent local yielding of the core. Bellfield
dam is discussed further in Section 7.10 and in Section 1.2 of Appendix G.

a) 0 b) 0

10 10

Concentrated zone of
Depth below crest (m) .

internal core settlement


Depth below crest (m) .

20 20

Concentrated zone of
internal core settlement

30 30

40 ES1 - Feb 1987 to Nov 1997 (21 to 31.5 years 40


post EOC)
ES1 - 21 years post EOC (to Feb 1987) ES1 - Feb 1987 to Nov 1997 (21 to 31.5
years post EOC)
EOC = end of embankment construction EOC = end of embankment construction
50 50
-100 0 100 200 300 400 500 0 10 20 30 40 50 60 70
Settlement (mm) Settlement (mm)

Figure 7.45: Post construction internal settlement of the core at Bellfield dam.

7.6.2 POST CONSTRUCTION TOTAL DEFORMATION OF SURFACE MONITORING POINTS

The total post construction surface deformations are presented as plots of settlement
versus dam height at selected times after the end of embankment construction. Plots of
the post construction lateral displacement during first filling are presented in Section
7.5.3. No additional lateral displacement plots are presented in this section.
Chapter 7: Deformation behaviour of embankment dams Page 7.103

Data is presented for the three embankment surface regions; the mid to downstream
crest, the mid to upper downstream slope, and the upper upstream to upstream crest
region (refer Figure 7.37). Base readings for the data are generally in the range from 0
to 0.5 years after the end of construction, most within 0.2 years, and include the period
of first filling for most dams.
The data are mainly for embankments on rock foundations. Several embankments
on soil/rock or soil foundations have been included where the foundation is considered
to have a limited influence on the dam settlement, or where it can be excluded by
deducting foundation settlement from base plate or base cross-arm readings from IVM
gauges.
The settlement data is presented in the following figures:
• Mid to downstream crest region - Figure 7.46 for 3 years, Figure 7.47 for 10 years
and Figure 7.48 for 20 to 25 years after the end of construction.
• Mid to upper downstream slope - Figure 7.49 for 3 years and Figure 7.50 for 10
years after the end of construction.
• Upper upstream slope to upstream crest region - Figure 7.51 for 3 years and Figure
7.52 for 10 years after the end of construction.

Plots for 20 to 25 years after the end of construction for the upstream and
downstream slope regions are presented in Section 2.1 of Appendix F.

(a) Mid to Downstream Crest Region


The data for the crest region is sorted based on core width, core material type and
placement moisture content. It can be seen from the figures that:
• The post construction crest settlements are generally much smaller than the core
settlement during construction (refer Figure 7.23).
• Nearly all dams experience less than 1% crest settlement post construction for
periods up to 20 to 25 years and longer after construction.
• Most experience less than 0.5% in the first 3 years and less than 0.75% after 20 to
25 years.

Table 7.19 summarises the typical range of crest settlement for various groupings of
core width, material type and placement moisture content. Embankments that are
clearly outliers (Ataturk, Djatiluhur, Beliche, Mita Hills, Roxo (near buttress), Rector
Chapter 7: Deformation behaviour of embankment dams Page 7.104

Creek dams) and potential outliers (Svartevann, Belle Fourche and Dixon Canyon
dams) have been excluded from this table. Smaller crest settlements (as a percentage of
the embankment height) are observed for dry placed clayey sands to clayey gravels
regardless of core width and dry to wet placed silty sands to silty gravels. A broader
range of crest settlement is shown for clay cores, wet placed clayey sand to clayey
gravel cores, and embankments with very broad core widths, most of which are dry
placed clays to sandy clays to clayey sands. For zoned earth and rockfill dams, poor
compaction of the rockfill is over-represented at the larger end of the range of crest
settlement.

Table 7.19: Typical range of post construction crest settlement.


Core Properties Crest Settlement (% of dam height) *1
No.
Core Width Moisture 20 to 25
Classn Cases 3 years 10 years
content years
Dry 9 0.05 to 0.55 0.10 to 0.65 0.20 to 0.95
CL/CH
Thin to Wet 11 0.04 to 0.75 0.08 to 0.95 0.20 to 1.10
medium Dry 5 0.10 to 0.25 0.10 to 0.40 < 0.5
SC/GC
Wet 18 0.15 to 0.80 0.20 to 1.10 < 1.1
Thin to Thick SM/GM All 16 0.06 to 0.30 0.10 to 0.65 < 0.5 to 0.7
CL/CH all (most dry) 12 0.02 to 0.75 0.10 to 1.0 0.5 to 1.0
Thick
SC/GC all (most dry) 5 0.05 to 0.20 0.10 to 0.35 0.10 to 0.45
Very Broad all all (most dry) 18 0.0 to 0.60 0.0 to 0.80 0.05 to 0.76
Note: *1 excludes possible outliers.

a) 6000
dry clay core, thin to medium width 3%
wet clay core, thin to medium width 2%
dry SC/GC core, thin to medium width
5000 Ataturk
wet SC/GC core, thin to medium width
dry SM/GM core, thin to medium width
Crest Settlement (mm) .

wet SM/GM core, thin to medium width


Thick clay cores
4000 Thick SC/GC cores
Thick SM/GM cores
Very broad cores

3000
1%

2000

1000

0
0 50 100 150 200 250 300
Note: Base readings range from 0 to 0.5
Dam Height (m)
years after embankment construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.105

b) 1400
1% 0.5%

Gepatsch
1200
Crest Settlement (mm) .

Djatiluhur
1000

Srinagarind

800 Blowering
Svartevann (left abut.)
dry clay core, thin to medium width
Beliche
Svartevann (main sect.) wet clay core, thin to medium width
600 dry SC/GC core, thin to medium width
Mita Hills
wet SC/GC core, thin to medium width
Rector Creek 0.25% dry SM/GM core, thin to medium width
400 wet SM/GM core, thin to medium width
Thick clay cores
Roxo
Thick SC/GC cores
200 Thick SM/GM cores
Very broad cores
poorly compacted rockfill
0
0 50 100 150 200 250 300
Note: Base readings range from 0 to 0.5
Dam Height (m)
years after embankment construction.

Figure 7.46: Post construction crest settlement at 3 years after end of construction, (a)
all data, (b) data excluding Ataturk.

1800
2% Gepatsch 1%
Djatiluhur
1600

1400
Crest Settlement (mm) .

0.5%

1200

Blowering
Rector Creek
1000

Beliche dry clay core, thin to medium width


800 wet clay core, thin to medium width
Mita Hills dry SC/GC core, thin to medium width
wet SC/GC core, thin to medium width
600 Roxo
dry SM/GM core, thin to medium width
wet SM/GM core, thin to medium width
0.25% Thick clay cores
400
Belle Fourche Thick SC/GC cores
Thick SM/GM cores
200 Very broad cores
poorly compacted rockfill
0
0 50 100 150 200 250 300
Note: Base readings range from 0 to 0.5
Dam Height (m)
years after embankment construction.

Figure 7.47: Post construction crest settlement at 10 years after end of construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.106

2500
Djatiluhur 1%

2%

2000
Crest Settlement (mm) .

1500

dry clay core, thin to medium width


Blowering wet clay core, thin to medium width
1000 dry SC/GC core, thin to medium width
Dixon Canyon wet SC/GC core, thin to medium width
Spring Canyon 0.5%
dry SM/GM core, thin to medium width
wet SM/GM core, thin to medium width
Thick clay cores
500 Thick SC/GC cores
0.25%
Thick SM/GM cores
Very broad cores
poorly compacted rockfill
0
0 50 100 150 200 250 300
Note: Base readings range from 0 to 0.5
Dam Height (m)
years after embankment construction.

Figure 7.48: Post construction crest settlement at 20 to 25 years after end of


construction.

(b) Mid to Upper Downstream Slope Region


For the mid to upper downstream slope region the data has been sorted based on
material type of the downstream shoulder fill. The embankments with rockfill
shoulders have been further divided into compaction rating of the rockfill.
Embankments with dry placed and poorly compacted rockfills have been highlighted.
The height designated for each case study is the height from the SMP to foundation
level.

(c) Upper Upstream Slope and Upstream Crest Region


For the upper upstream slope and upstream crest region the data has been sorted based
on material type of the upstream shoulder fill. The embankments with rockfill
shoulders have been further divided into compaction rating of the rockfill.
Embankments with dry, poorly compacted rockfills in the upstream shoulder have been
highlighted.
The height designated for each case study is the height from the SMP to foundation
level.
Several zoned earthfill embankments with very broad cores where the foundation
potentially has had an influence on the settlement have been included in the plots for the
upstream and downstream slope. For these cases the height used is that from the SMP
Chapter 7: Deformation behaviour of embankment dams Page 7.107

to bedrock foundation. Excluded though were cases where the foundation contributed
significantly to the settlement such as at Fresno dam and the downstream slope of
Horsetooth dam.

1200
2% 1% Rockfill - well-compacted
Rockfill - reasonably to well compacted
Eildon
Rockfill - reasonably compacted
1000 Rockfill - poorly compacted
Svartevann gravel shoulders
earthfill shoulders and dams with very broad cores
Srinagarind Dry, poorly compacted rockfills
Settlement (mm) .

800

0.5%
600
Eppalock 0.25%
Gepatsch

400
Pueblo Beliche
(left)

200

0
0 50 100 150 200 250

Height from SMP to foundation level (m)


Figure 7.49: Post construction settlement of the downstream slope (mid to upper region)
at 3 years after end of construction.

1500
2%
Rockfill - well-compacted
Eildon Rockfill - reasonably to well compacted
Rockfill - reasonably compacted
1200 1% Rockfill - poorly compacted
gravel shoulders
earthfill shoulders and dams with very broad cores
Settlement (mm) .

Dry, poorly compacted rockfills


Ataturk
900 0.5%

Eppalock
Gepatsch
600 0.25%
Beliche
Pueblo
(left)

300

0
0 50 100 150 200 250

Height from SMP to foundation level (m)


Figure 7.50: Post construction settlement of the downstream slope (mid to upper region)
at 10 years after end of construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.108

1200
1% 0.5%

Gepatsch

900
Eildon Djatiluhur
Cougar
Settlement (mm) .

Netzahualcoyotl
Beliche 0.25%

Cherry Valley
600
Glenbawn
Eppalock

Rockfill - well-compacted
Rockfill - reasonably to well compacted
300 Rockfill - reasonably compacted
Rockfill - poorly compacted
gravel shoulders
earthfill shoulders and dams with very broad cores
Dry, poorly compacted rockfills
0
0 50 100 150 200 250 300
Height from SMP to foundation level (m)

Figure 7.51: Post construction settlement of the upper upstream slope and upstream
crest region at 3 years after end of construction.

1500
2% 1% 0.5%

Djatiluhur
1200

Beliche
Eildon
Settlement (mm) .

900
Spring Canyon Dixon Canyon

Horsetooth 0.25%

600 Eppalock

Rockfill - well-compacted
Rockfill - reasonably to well compacted
Rockfill - reasonably compacted
300 Rockfill - poorly compacted
gravel shoulders
earthfill shoulders and dams with very broad cores
Dry, poorly compacted rockfills
0
0 50 100 150 200 250 300
Height from SMP to foundation level (m)
Figure 7.52: Post construction settlement of the upper upstream slope and upstream
crest region at 10 years after end of construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.109

Table 7.20: Typical range of post construction settlement of the upstream and
downstream shoulders.
Downstream Shoulder *1, *2 Upstream Shoulder *1, *2
Material Compaction Settlement Settlement
Type Rating (% H from SMP to fndn) (% H from SMP to fndn)
No. 3 years 10 years No. 3 years 10 years
well 11 0.05 to 0.35 0.05 to 0.55 12 0.10 to 0.60 0.10 to 0.70
reas to well 9 < 0.30 < 0.50 11 0 to 0.55 0.10 to 0.60
Rockfill reas 5 0.20 to 1.0 0.10 to 1.0 3 < 0.70 - *4
poor 4 0.10 to ? * 4 0.15 to ? * 4 8 0.10 to 1.05 0.15 to 1.20
poor – dry *3 7 0.15 to 1.60 0.30 to 2.00 5 0.15 to 1.35 0.20 to 1.6
Gravels - 7 < 0.15 < 0.25 3 < 0.15 < 0.25
Earthfills - 20 0.0 to 0.40 0.0 to 0.70 14 0.05 to 0.60 0.10 to 0.70
Note: *1 Excludes possible outliers.
*2 Settlements quoted are a percentage of the height from the SMP to foundation level.
*3 For the dry placed poorly compacted rockfills, settlements at Upper Yarra, Matahina and El
Infiernillo are much less than for the other cases, for which settlements are close to or greater
than 1% of the height.
4
* insufficient data.

From the figures and as summarised in Table 7.20 it can be seen that for the
embankment shoulders:
• Settlements of up 1 to 2% are observed for poorly compacted rockfills, both in the
upstream and downstream shoulder. Greater settlements are observed for the dry
placed, poorly compacted rockfills.
• For reasonably compacted rockfills the range of settlement is quite broad, from
0.1% up to 1.0%, for a limited number of cases. At Svartevann and Gepatsch dams
the greater settlements are attributable to large collapse type settlement of dry placed
rockfills, and at Ataturk dam to collapse settlement of weathered rockfills.
• Much lower settlements, generally less than 0.5 to 0.7% at ten years after
construction, are observed for well and reasonably to well compacted rockfills, and
for embankments with compacted earthfill in the shoulder.
• Very low settlements (less than 0.25% at 10 years) are observed for embankments
with gravel shoulders.

Possible outliers include Svartevann, Horsetooth (upstream slope), Dixon Canyon,


Spring Canyon, Srinagarind and the downstream slope of the left abutment embankment
Chapter 7: Deformation behaviour of embankment dams Page 7.110

at Pueblo dam. Several of these cases are discussed further in Section 7.10 and 7.11,
and in Appendix G.

7.6.3 POST CONSTRUCTION CREST SETTLEMENT VERSUS TIME

The post construction crest settlement versus time for the case studies are presented in
Figure 7.53 to Figure 7.61 in the form of settlement as a percentage of the embankment
height versus time (in years since end of construction) on a log scale. The case studies
have been sorted based on core width, core material type and placement moisture
content into the following figures:
• Figure 7.53 – thin to medium core widths, dry placed clay cores;
• Figure 7.54 – thin to medium core widths, wet placed clay cores;
• Figure 7.55 – thin to medium core widths, dry placed clayey sand to clayey gravel
cores;
• Figure 7.56 – thin to medium core widths, wet placed clayey sand to clayey gravel
cores;
• Figure 7.57 – thin to medium core widths, silty sand to silty gravel cores both dry
and wet placed;
• Figure 7.58 – clay cores of thick core width;
• Figure 7.59 – silty and clayey sand and gravel cores of thick core width;
• Figure 7.60 – very broad cores where the foundation has limited influence on
settlement; and
• Figure 7.61 – very broad cores where the foundation has or potentially has a
significant influence on settlement. This figure is plotted with the total settlement in
millimetres.

The foundation influence has been evaluated from IVM or foundation base plate
records where available. However, this was not possible at Pueblo, Dixon Canyon and
Spring Canyon dams because foundation settlements were not measured post
construction. For Horsetooth and Steinaker dams, the embankment only settlement
from IVM records is included in Figure 7.60 and the combined embankment and
foundation settlement from SMP records is included in Figure 7.61.
In most cases the base survey reading was within 0.5 years of end of embankment
construction, but for a number of cases it was after this time. In several cases first
Chapter 7: Deformation behaviour of embankment dams Page 7.111

filling had been completed prior to the start of monitoring, these cases included but are
not limited to, Nillahcootie, Peter Faust and Eildon dams.
The size of each figure has been set to the same margins for each plot area and the
time interval standardised to cover from 0.1 to 100 years to allow for visual comparison
between the figures. In most cases a standard vertical axis of 0 to 1.2% settlement has
been used, but for several plots where settlements were larger a broader vertical axis has
been adopted. If the case studies represented in a particular plot include an outlier/s
(such as Ataturk in Figure 7.53a) a secondary plot excluding the outlier/s is presented as
figure b.
Several other points to note from the figures are:
• Where the data does not extend to the y-axis it is because the base reading is or has
assumed to be at the end of construction; i.e. zero time.
• The end of first filling has been indicated in the figures by an arrow for each case
study. There are several reasons why first filling is not indicated for a number of
cases studies:
− First filling was not completed in the period of deformation shown (e.g. Ataturk
and La Angostura dams).
− First filling was completed prior to the base survey reading (e.g. Nillahcootie,
Peter Faust and Eildon dams).
− For cases with the base reading at time = 0 years, first filling occurred in the
time period up to the first reading point after the base survey (e.g. Talbingo and
Bellfield dams).
− The time of the end of first filling is not known (e.g. Tooma dam).

The assessment of “wet” or “dry” placement has been based on the average moisture
content at placement in relation to Standard optimum moisture content and/or the pore
water response during construction from piezometers installed in the core. The
following criteria were used:
• “Wet” placement for cores placed from slightly dry (0.2 to 0.3% dry) to wet of
Standard optimum moisture content and/or where the pore water pressure response
indicated positive pore water pressure coefficients that exceeded about 0.1 to 0.2 at
end of construction. As previously discussed (Section 7.4.1.4) earthfills placed
wetter than about 0.5% dry of Standard optimum tend to develop positive pore water
Chapter 7: Deformation behaviour of embankment dams Page 7.112

pressures during construction, although this is material type and stress level
dependent.
• “Dry” placement for cores placed on average drier than about 0.5% dry of Standard
optimum moisture content or where positive pore water pressures developed during
construction indicated a pore water pressure coefficient less than about 0.1 to 0.2.

For a number of cases no details were available on the moisture content at


placement or the pore water pressure data during construction if piezometers had been
installed. For these cases a judgement was made as to whether the core was likely to
have been placed “dry” or “wet”. Some of the judgements made were:
• For the Japanese dams Seto, Kurokawa, Shimokotori, Taisetsu and Kisenyama
dams, all central core earth and rockfill dams with thin to medium core widths,
limited information was available on the core materials. Most of these dams were
constructed in the 1970’s and were assumed to be predominantly clayey gravels and
wet placed given material availability and typical procedures used in Japan for
embankment construction at the time (Takahashi and Nakayama 1973; Shiraiwa and
Takahashi 1985; Kanbayashi et al 1979).
• For embankments with silty sand or silty gravel cores no distinction was made with
respect to “wet” or “dry” placement for the post construction deformation. Where
no information was available it was assumed that construction pore water pressures,
if developed, had been dissipated to small values a short time period after end of
construction. Embankments with limited or no data on placement moisture
condition or pore water pressure response included Mud Mountain, Round Butte,
Cherry Valley and Mammoth Pool dams.
• For the Tennessee Valley Authority dams Nottely, Watuaga and South Holston, wet
placement of the clay cores was assumed as implied by Leonard and Raine (1958)
and Blee and Riegel (1951).
• For Googong, Spilt Yard Creek, Peter Faust, Bjelke Peterson, Corin and Benmore
dams the classification is likely to be borderline between “wet” and “dry” as
placement was or was likely to have been within about 0.5% of Standard optimum
moisture content. All were classified as “dry” placed.
• Cairn Curran, a zoned earthfill embankment with very broad earthfill core, “dry”
placement was assumed.
Chapter 7: Deformation behaviour of embankment dams Page 7.113

a) 0.0% 6 9
5
4 3
0.5%
7
8 2
1.0%
Settlement (% of dam height) .

1.5%

2.0%

2.5%

3.0%

3.5%

4.0% 1

4.5%
0.1 1.0 10.0 100.0

Time (years since end of construction)


1. Ataturk (960) 2. Bellfield (704) 3. Matahina (4) 4. Split Yard Creek (21) 5. Nillacootie (701)
6. Peter Faust (15) 7. Windemere (31) 8. Glenbawn Saddle (19) 9. Bjelke Petersen (13)

b) 0.0%

9
6 6
9 5
0.2%
Settlement (% of dam height) .

4
0.4%
3

7
0.6%
8

2
0.8%

End of first filling


1.0%

1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

Figure 7.53: Crest settlement versus time for zoned embankments with thin to medium
width central core zones of dry placed clayey earthfills; (a) all data, (b) data excluding
Ataturk.
Chapter 7: Deformation behaviour of embankment dams Page 7.114

0.0%

8
7
7 3 4
0.5% 9
3
Settlement (% of dam height) .

1.0%

1.5%

2.0%

End of first filling

1
2.5%
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Djatilihur 2. El Infiernillo (M10) 3. Wivenhoe (SS26) 4. Talbingo (CS6) 5. Blowering (IVM A)


6. Canales 7. William Hovell (Ch. 503) 8. La Agnostura (Pnt. d) 9. Netzahualcoyotl (14-1u)

Figure 7.54: Crest settlement versus time for zoned embankments with thin to medium
width central core zones of wet placed clayey earthfills.

0.0%

0.2%
5
Settlement (% of dam height) .

4
2
0.4%

1
0.6%

1. Copeton (74)
0.8% 2. Googong (S6)
3. Corin (24)
4. Wyangala (32)
5. Fukada

1.0%

End of first filling


1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

Figure 7.55: Crest settlement versus time for zoned embankments with thin to medium
width central core zones of dry placed clayey sand to clayey gravel (SC to GC)
earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.115

a) 0.0%
4

0.2%
6
Crest settlement (% of dam height) .

5
0.4% 9
1 1 4
5 8
0.6% 8
3
7

0.8%

6
1.0%
2

1.2%

1.4%
End of first filling
10
1.6%
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Bath County Upper Dam 2. Gepatsch (3) 3. Srinagarind 5. Shimokotori 6. Seto
4. Upper Dam, Korea (USP4) 7. Dartmouth (CS7) 8. Thomson (CS6) 9. Chicoasen (I-AB) 10. Beliche (J54)

b) 0.0%
18

18
17
0.2% 11
Crest settlement (% of dam height) .

16
13
12 17 12
13 15
0.4%
16
14

0.6%

0.8%

1.0%

1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

11. Tokachi 12. Terauchi 13. Kurokawa 14. Tedorigawa


15. Taisetsu 16. Kisenyama 17. Rosshaupten 18. Soyang (SP6)

Figure 7.56: Crest settlement versus time for zoned embankments with thin to medium
width central core zones of wet placed clayey sand to clayey gravel (SC to GC)
earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.116

7
0.0%

6 1
6
0.2%
7 5
1
Crest settlement (% of dam height) .

9 3
5 2
0.4%

0.6%
8

0.8%

1.0%
10
End of first filling

1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Round Butte (19) 2. Andong (CS7) 3. Ajuare 4. Svartevann (main section) 5. Parangana (Ch. 400)
6. Geehi (5) 7. Rowallan (Ch. 1500) 8. Mud Mountain 9. Cougar (TD-11) 10. Svartevann (left abut.)

Figure 7.57: Crest settlement versus time for zoned embankments with thin to medium
width central core zones of silty sand to silty gravel (SM to GM) earthfills.

1 6 12
0.0%

5 1
12
8
0.2%
Crest settlement (% of dam height) .

0.4% 7
6 8

4 9
0.6%
10
9
5
0.8%
3

11
1.0% 2

End of first filling


7
1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Waipapa (DM16) 2. South Holston 3. Watuaga 4. Nottely 5. Upper Yarra (M5) 6. Maroon (25)
7. Eppalock (CS2) 8. Burrendong (73) 9. Navajo (Stn 38) 10. Eildon (CS7) 11. Glenbawn (G2H) 12. San Justo (Stn 3.3)

Figure 7.58: Crest settlement versus time for zoned embankments with central core
zones of clayey earthfills of thick width (1 to 2.5H to 1V combined width).
Chapter 7: Deformation behaviour of embankment dams Page 7.117

0.0% 2
3
6
4
2
7
0.2% 4 8
Crest settlement (% of dam height) .

5
8 6

7 1 3
1
9
0.4% 9

0.6%

0.8%

1.0%

End of first filling


1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Cardinia (SS13) 2. Greenvale (CS5) 3. Bradbury (Stn 19) 4. Benmore (DM16) 5. Tooma (IVM A)
6. Pukaki (DM11) 7. Cherry Valley (Stn 20) 8. Mammoth Pool (B11) 9. Meeks Cabin (Stn 16.5)

Figure 7.59: Crest settlement versus time for zoned embankments with central core
zones of silty to clayey gravel and sand (SC, GC, SM, GM) earthfills of thick width (1
to 2.5H to 1V combined width).

a) 0.0%

7
1, 2
0.4%
Crest settlement (% of dam height) .

2 10
0.8%
10 1
8

1.2%
8

5
1.6%
End of first filling 9

6 3
2.0%
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Pueblo, right emb. (Stn 46) 2. Pueblo, left emb. (Stn 77.5) 3. Belle Fourche (Stn 31) 4. Belle Fourche (MP10) 5. Mita Hills (A4)
6. Roxo (20m from buttress) 7. Roxo (95m from buttress) 8. Horsetooth (IVM A) 9. Rector Creek (Stn 16) 10. Steinaker (IVM A)
Chapter 7: Deformation behaviour of embankment dams Page 7.118

b) 0.0%

15
16
Crest settlement (% of dam height) .

0.2%
17
14 17
14
11, 12 16 10 13
0.4% 11
12

0.6%

0.8%

1.0%
End of first filling

1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)
10. Tullaroop (CS2) 11. San Luis (Stn 110) 12. San Luis, slide area (Stn 138) 13. Cairn Curran (C3)
14. Biet Netufa (IVM) 15. Eucembene (15) 16. Khancoban (12/13) 17. Enders (Stn 48.4)

Figure 7.60: Crest settlement versus time for embankments with very broad core widths
(> 2.5H to 1V combined width) and limited foundation influence.

0
1 to 4

200
Crest settlement (mm) .

400

3
5
600
6
6 1

800
7

1000
End of first filling

2
1200
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Horsetooth (Stn 10) 2. Dixon Canyon (Stn 9) 3. Soldier Canyon (Stn 11) 4. Spring Canyon (Stn 5)
5. San Luis (Stn 86) 6. Steinaker (Stn 14.6) 7. Fresno (Stn 21)

Figure 7.61: Crest settlement versus time for embankments with very broad core widths
(> 2.5H to 1V combined width) and potentially significant foundation influence.
Chapter 7: Deformation behaviour of embankment dams Page 7.119

The typical trend of post construction crest settlement versus log time is for a near linear
relationship for most case studies, particularly after the end of first filling. Although, a
gradual increase in the crest settlement rate (rate per log cycle of time) with time, or in-
fact a decrease in rate with time is not unusual.
Before getting into a detailed discussion on the post construction crest settlement
behaviour, the data on the long-term or post first filling crest settlement rate (rate per
log cycle of time) is presented. Figure 7.62 presents the long-term crest settlement rate
versus embankment height for zoned earthfill and earth-rockfill embankments with
cores of thin to thick widths, and Figure 7.63 for earthfill and zoned embankments with
very broad core widths. The data has been sorted based on reservoir operation. For the
embankments of thin to thick core width the data has also been sorted based on core
material type and those embankments with poorly compacted upstream rockfill zones
have been highlighted. Table 7.21 presents a summary of the typical range of crest
settlement rate for various sub-groupings of the case study data.
The assessment of reservoir operation post first filling is:
• “Fluctuating” reservoir where the reservoir is subject to a seasonal (usually annual)
or regular (more than once per year) drawdown that is typically greater than about
0.1 to 0.15 times the height of the embankment at its maximum section.
• “Steady” or “slow” reservoir operation for embankments where the reservoir
remains steady over time, is subject to fluctuations that occur slowly over time (i.e.
slow drawdown over several years) or where the seasonal or regular fluctuation is
less than 0.1 to 0.15 times the height of the embankment at its maximum section.

The crest settlement rate for each case study has been determined from the
settlement versus log time plots in the above figures from the portion of the curve after
the end of first filling. Where the settlement rate increases with time, such as for Enders
dam in Figure 7.60b, the longer term rate has been used. For case studies where the
settlement post first filling includes short periods of rapid acceleration, such as at
Eppalock, Eildon and Djatiluhur dams, the representative rate for these embankments
excludes the periods of rapid acceleration.
Chapter 7: Deformation behaviour of embankment dams Page 7.120

a) 3.5
Clayey core, fluctuating reservoir
Clayey core, steady or slow reservoir Ataturk
3 SC/GC core, fluctuating reservoir
Long-term Crest Settlement Rate .

SC/GC core, steady or slow reservoir


SM/GM core, fluctuating reservoir
SM/GM core, steady or slow reservoir
(% per log time cycle) .

2.5

1.5

0.5

0
0 25 50 75 100 125 150 175 200
Embankment Height (m)
b) 1.25
Indicates poorly compacted
Djatilihur
upstream rockfill
Long-term Crest Settlement Rate .

1
Eppalock
(% per log time cycle) .

Svartevann (left abut.)


Gepatsch
Upper Dam
0.75

Bellfield
0.5
Maroon

Windemere

0.25

0
0 25 50 75 100 125 150 175 200
Embankment Height (m)
Clayey core, fluctuating reservoir SC/GC core, fluctuating reservoir SM/GM core, fluctuating reservoir
Clayey core, steady or slow reservoir SC/GC core, steady or slow reservoir SM/GM core, steady or slow reservoir

Figure 7.62: Long-term post construction crest settlement rate for zoned earthfill and
earth-rockfill embankments of thin to thick core widths, (a) all data, and (b) data
excluding Ataturk.
Chapter 7: Deformation behaviour of embankment dams Page 7.121

a) 5
Long-term Crest Settlement Rate . Belle Fourche (closure section) Steady or slow reservoir
(75 to 85 yrs) Fluctuating reservoir
4
(% per log time cycle) .

Roxo (near buttress)

2
Belle Fourche Rector Creek
(first 10 to 12 yrs)

0
0 25 50 75 100 125 150 175 200
Embankment Height (m)

b) 1.5

Dixon Canyon Steady or slow reservoir


Long-term Crest Settlement Rate .

Spring Canyon Fluctuating reservoir


1.25
(% per log time cycle) .

0.75
Horsetooth
Pueblo (left)
Soldier Canyon
Pueblo (right)
0.5

Enders

0.25

0
0 25 50 75 100 125 150 175 200
Embankment Height (m)
Figure 7.63: Long-term post construction crest settlement rate for embankments of very
broad core width, (a) all data, and (b) data excluding Belle Fourche, Roxo and Rector
Creek.

As a whole the total settlement, settlement versus time and settlement rate plots provide
a comprehensive summary of the post construction crest settlement behaviour for
earthfill and earth-rockfill embankments. The data allows for comparisons of
magnitude and rate based on core width, core material type and moisture content at
placement as well as consideration of reservoir operation post first filling and the use of
poorly compacted rockfills in the upstream shoulder. Table 7.19 and Table 7.21 provide
typical ranges of total crest settlement and long-term crest settlement rate sorted for
most of these factors.
Chapter 7: Deformation behaviour of embankment dams Page 7.122

Clear and possible outliers to the general trend of post construction crest settlement
are generally evident in at least two of the plot types, usually the settlement versus time
plot and then either or both of the total settlement and settlement rate plots. The
settlement versus time plots are the most comprehensive for assessment of potential
outliers. In addition to highlighting large total settlement or high settlement rates, other
trends that are potentially indicative of “abnormal” deformation behaviour are evident,
such as acceleration in settlement over short time periods. Embankments indicating this
type of behaviour include Djatiluhur, Canales, El Infiernillo, Beliche, Eppalock, Eildon,
Upper Yarra, Svartevann, Matahina and possibly Blowering and Wyangala dams. A
number of these case studies are discussed in Section 7.10. Often, but not always, these
accelerations in settlement rate are associated with drawdown (refer to point d below).
It is important to note that use of the term “abnormal” in relation to the deformation
behaviour of an embankment does not necessarily equate with potential instability. In
nearly all cases identified as “abnormal” or potentially “abnormal” the overall stability
of the embankment is not in question.

Table 7.21: Summary of long-term crest settlement rates (% per log cycle of time).
Core Core Reservoir No. Long-term *1
Comments
Width Classn Operation Cases Settlement Rate
Fluctuating 14 0.09 to 0.57 Higher rates at Djatiluhur (1.17%) and
Eppalock (0.91%).
CL/CH Steady or slow 15 0.04 to 0.50 Very high at Ataturk (3.1%).
Higher rates at Maroon (0.46%),
(most < 0.26)
Windemere (0.36%) and Bellfield
(0.50%).
Thin to
Fluctuating 18 0.06 to 0.37 Higher rates at Upper Dam (0.75%) and
thick SC/GC Gepatsch (0.83%).
Steady or slow 8 0 to 0.26
Fluctuating 12 0.03 to 0.21 Higher rates at Svartevann (0.87%) and
Andong (0.39%).
SM/GM
Steady or slow 2 < 0.10 Only two cases, Round Butte &
Parangana
Very high rates (> 1.8%) at Belle Fourche
0.07 to 0.70 & Roxo (near buttress).
All types Fluctuating 17 High rates at Rector Creek (1.65%),
Very broad (most < 0.35) Dixon Canyon (1.38%) and Spring
(most CL) Canyon (1.31%).
Steady or slow 2 0.08 and 0.44 Only two cases, Eucembene (0.08%) and
Enders (0.44%).
*1 Long-term settlement rate in units of settlement as a percentage of the embankment height per log
cycle of time.
Chapter 7: Deformation behaviour of embankment dams Page 7.123

The factors affecting the post construction settlement of the mid to downstream region
of the crest are:

(a) Influence of core material type and core width on the post construction crest
settlement
Core material type has an influence on the post construction crest settlement. In
general:
• Earthfill cores of silty sands to silty gravels typically show lower total crest
settlements (Table 7.19) and lower long-term settlement rates (Table 7.21).
• Clay earthfill cores, on average, tend to have high long-term crest settlement rates
(Figure 7.62 and Table 7.21) with clayey sand to clayey gravel cores showing
intermediate rates.

Core width also influences the post construction crest settlement behaviour, but its
influence is only evident for the crest settlement during and shortly after first filling.
During first filling, crest settlements for most of the case studies with very broad
earthfill cores (Figure 7.60 and Figure 7.61) and thick earthfill cores (Figure 7.58 and
Figure 7.59) are limited, while those for thin to medium width cores are larger. This is
likely to be due to several factors:
• The direction of the water load on first filling acting on the upstream face of the core
and its distance from the dam axis.
• For embankments with thick cores, the core is generally not reliant on the shoulders
zones for support and consequently not significantly influenced if collapse
settlement on first filling occurs in the upstream shoulder. Conversely,
embankments of thin to medium core width are more reliant on the shoulders for
support and therefore can be significantly influenced by collapse settlements in the
upstream shoulder on first filling.

The core width appears to have no recognisable influence on the long-term post first
filling crest settlement rate (Table 7.21, Figure 7.62 and Figure 7.63). For
embankments with very broad core width the long-term crest settlement rates (Figure
7.63) are highly variable and are discussed further in (f) below.
Chapter 7: Deformation behaviour of embankment dams Page 7.124

(b) Influence of compaction moisture content on the post construction crest


settlement
For embankments with cores of thin to medium width it is evident that, after excluding
the case studies likely affected by collapse settlement of the upstream rockfill, the range
of crest settlement at any given time for cores placed wet are comparable to those for
dry placed cores. For example, for the cases where monitoring started close to zero
time the crest settlement at 10 years after end of construction are in the range:
• 0.2 to 0.65% for dry placed clay earthfills (Figure 7.53).
• 0.15 to 0.7% for wet placed clay earthfills (Figure 7.54).
• 0.1 to 0.4% for dry placed clayey sand to clayey gravel earthfills (Figure 7.55).
• 0.2 to 0.5% for wet placed clayey sand to clayey gravel earthfills (Figure 7.56).

This is not to say that compaction moisture content has a negligible influence on the
post construction settlement behaviour, but that other factors have a greater influence
and over-shadow that of compaction moisture content such as reservoir operation and
collapse settlement of the upstream rockfill for wet placed cores.
The effect of compaction moisture content on the post construction internal vertical
strain is clearly evident at Talbingo dam. In the lower region of the wet placed core at
Talbingo dam reductions in pore water pressure post first filling were very large, up to
600 to 1100 kPa, and post construction vertical strains were up to 2.5 times greater tha n
in the mid to upper region where the core was placed drier and reductions in pore water
pressures were consequently much smaller (Figure 7.43 in Section 7.6.1). This
reduction in pore water pressure in the core at Talbingo dam was much greater than that
at Blowering, Thomson and Dartmouth dams (dams all greater than 100 m height) yet
the post construction crest settlement at Talbingo is less than for these other
embankments.
The limited influence of compaction moisture content on post construction crest
settlement can also be attributable to the usually small reduction in pore water pressure
from end of construction to close to equilibrium conditions for wet placed clay cores. In
some cases, such as at Talbingo dam, where the reduction in pore water pressure is large
(more than 200 to 500 kPa) it is usually confined to small regions of the core.
Therefore, the net influence of consolidation due to dissipation of pore water pressures
on the overall crest settlement will be relatively small.
Chapter 7: Deformation behaviour of embankment dams Page 7.125

Table 7.19 does indicate that a higher range of crest settlement is observed for wet
placed clay and clayey sand to gravel cores of thin to medium width. However, the
range of settlement for most of these case studies with wet placed cores is similar in
magnitude to that of dry placed cores, as shown in Figure 7.53 to Figure 7.56. The
reason for the higher range of settlement is that the data for wet placed cores does
includes several case studies where the crest settlement was thought to be strongly
influenced by collapse compression in the upstream rockfill.

(c) Influence of rockfill placement procedures for zoned embankments with


rockfill in the upstream shoulder
The data for zoned earth and rockfill embankments shows that case studies with poorly
compacted rockfill in the upstream shoulder are over-represented at the higher end of
the total settlement plots (Figure 7.46 to Figure 7.48) and the higher end of long-term
settlement rate plot (Figure 7.62). A more correct conclusion from the data would be
that embankments where the rockfill in the upstream shoulder is susceptible to large
magnitude deformations due to collapse compression on wetting represent a very large
portion of the data at the higher end of these plots.
Earlier findings (Section 7.5.2) indicate that the magnitude of deformation on
collapse compression is related to a number of factors other than degree of compaction,
including placement moisture content, rock substance unconfined compressive strength
and its reduction on saturation, stress level and particle size distribution. Assuming the
magnitude of deformation of the upstream rockfill due to collapse compression is large
on first filling, the shear strength properties of the core strongly influence the magnitude
of crest settlement, with wet placed cores of low undrained strength shown to be
susceptible to large crest settlements on first filling. The post construction settlement
data tends to confirm this, although large settlements are not just confined to wet placed
cores, they are also observed for dry placed cores. This is discussed further in Section
7.10.
From Figure 7.46, the plot of total settlement at 3 years after construction, crest
settlements greater than 0.35 to 0.4% of the embankment height occur in embankments
where the magnitude of settlement of the upstream rockfill due to collapse settlement on
wetting was significant, including:
• A number of embankments with poorly compacted rockfill in the upstream shoulder.
Chapter 7: Deformation behaviour of embankment dams Page 7.126

• Embankments with dry placed and reasonably compacted (compacted in 2 m lifts)


rockfill in the upstream shoulder (Gepatsch and Svartevann dams).
• Embankments with wetted and reasonably compacted weathered rockfill in the
upstream shoulder (Ataturk and Beliche dams).
• Blowering dam, where the well-compacted rockfill was susceptible to large loss in
strength on wetting.

The greater total settlement in these embankments is also reflected in the plots at 10
years and 20 to 25 years after construction.
The long-term crest settlement rate data (Figure 7.62) shows that for almost all cases
where the settlement rate is greater than about 0.4% per log cycle of time the upstream
rockfill (and usually the downstream rockfill as well) was poorly compacted and/or
susceptible to large magnitude deformations due to collapse compression. This is an
important observation because the influence of rockfills susceptible to collapse
compression on the embankment deformation behaviour is generally only related to the
embankment performance during and shortly after first filling. But the high long-term
settlement rates for these embankments suggests that they are more susceptible to
softening effects and degradation long-term as a consequence of factors including poor
lateral support of the core, lateral spreading, internal cracking and softening in
undrained strength. This is discussed further in Section 7.10 referencing the
deformation behaviour from selected case studies.

(d) Influence of reservoir operation on the long-term crest settlement rate


The reservoir operation has a significant influence on the long-term post first filling
crest settlement rate for zoned embankments as shown in Figure 7.62 and summarised
in Table 7.21. Most of the periods of rapid acceleration of the crest settlement after first
filling are associated with drawdown events. Exceptions to this are the earthquake
influence at El Infiernillo and Matahina dams.
The difference in crest settlement behaviour due to reservoir operation is highlighted
in Figure 7.64 and Figure 7.65. Both figures are of central core earth and rockfill
embankments with clay cores of thin to thick width, Figure 7.64 for embankments
where the reservoir is relatively steady and Figure 7.65 for embankments where the
reservoir undergoes seasonal (often annual) drawdown.
Chapter 7: Deformation behaviour of embankment dams Page 7.127

0.0%
1
1
4
0.2%
Crest Settlement (% of dam height) .

3 influence of
earthquake
2
2 3
0.4%
4

1. Waipapa (DM16)
0.6% 2. Wivenhoe (SS26)
3. Talbingo (CS6)
4. Matahina (4)

0.8%

End of first filling

1.0%

1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

Figure 7.64: Crest settlement versus time for central core earth and rockfill
embankments with central core zones of clayey earthfills and steady post first filling
reservoir operation.

0.0%
2
5 4 3

0.2%
Crest settlement (% of dam height) .

0.4%

0.6%
5
1
3
0.8%

1.0%
End of first filling
4

1
1.2%
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Blowering (IVM A) 2. William Hovell (SS26) 3. Upper Yarra (M5) 4. Eppalock (CS2) 5. Eildon (CS7)

Figure 7.65: Crest settlement versus time for central core earth and rockfill
embankments with thin to thick central core zones of clayey earthfills subjected to
seasonal reservoir fluctuation.
Chapter 7: Deformation behaviour of embankment dams Page 7.128

For the seasonal drawdown case studies the rockfill placement consisted of:
• Placed dry in 1.5 to 2 m layers and spread with dozer tracking for Eildon and
Eppalock, and possibly Upper Yarra dams.
• Wetted and well-compacted for Blowering dam, but the rockfill is susceptible to
large strength loss on wetting.
• Reasonably to well compacted for William Hovell.

For Eildon, Eppalock, Upper Yarra and Blowering dams the periods of acceleration
of the crest settlement post construction (Figure 7.65) are associated with large
drawdown events. Other case studies where acceleration in crest settlement occurs on
drawdown include Djatiluhur, Beliche and Wyangala dams. These cases are discussed
further in Section 7.10 and in Appendix G.

(e) Crest settlement during first filling


Higher rates of crest settlement (per log cycle of time) during first filling are generally
observed for zoned embankments of thin to medium core width, but only some of these
embankments show this behaviour (Figure 7.53 to Figure 7.57). Dams with higher rates
of settlement during first filling include:
• Split Yard Creek and Glenbawn Saddle dams, both of dry placed clay cores of thin
to medium width.
• Blowering, Canales and Netzahualcoyotl dams, all wet placed clay cores of thin to
medium width.
• Googong dam, a dry placed clayey sand core of medium width.
• Thomson, Beliche, Gepatsch and Dartmouth dams, all wet placed clayey sand to
clayey gravel cores of thin to medium width.
• Svartevann, Cougar and Rowallan dams, all wet placed silty sand to silty gravel
cores of thin to medium width.
• Nottely and South Holston dams, both clay cores of thick width and possibly wet
placed.
• Mita Hills and Horsetooth dams, embankments with very broad earthfill zones.

Apart from Mita Hills and Horsetooth dams, all embankments are central core earth
and rockfill embankments.
Chapter 7: Deformation behaviour of embankment dams Page 7.129

It is suspected that the mechanism of deformation for most of the case studies
includes collapse settlement of the upstream rockfill on first filling and resultant plastic
deformation with potential development of localised shearing within the core. This
mechanism was discussed in Section 7.5.2 with the aid of finite difference modelling
and is discussed in more detail for selected case studies in Section 7.10 and Appendix
G.

(f) Influence of very broad core width


For earthfill embankments or zoned embankments with very broad cores, notable
aspects of the post construction crest settlement behaviour are:
• For a number of case studies the settlement versus time plots (Figure 7.60 and
Figure 7.61) show an increasing rate of settlement after first filling. This trend is
clearly evident for Enders dam and the Horsetooth Reservoir dams (Horsetooth,
Dixon Canyon, Spring Canyon and Soldier Canyon dams), and to a lesser degree for
San Luis dam and several other embankments.
• The long-term settlement rate of these embankments shows a high degree of
variability (Figure 7.63), ranging from 0.07% to 4.5% within the closure section at
Belle Fourche dam many years after construction.

With respect to the long-term settlement rate, the very high rates (> 1.6% per log
time cycle) at Belle Fourche (within the closure section), Rector Creek, Mita Hills and
Roxo dam (adjacent to the buttress) are considered “abnormally” high. Within the
closure section at Belle Fourche dam it is notable that the settlement rate at 75 to 85
years after construction is almost 3 times what is was at 10 to 12 years after
construction. The high rates at Dixon Canyon (1.38%) and Spring Canyon (1.31%)
dams are also considered to be “abnormally” high. For these two dams there is a query
over the influence of the foundation on the long-term crest settlement, but from review
of the data records (refer Section 3.3 of Appendix G) this influence is considered to be
limited and most of the long-term crest settlement is attributable to deformation within
the earthfill.
The remaining embankments are considered to represent “normal” long-term
settlement rates with values ranging from 0.07% to 0.7% per log time cycle.
Embankment height appears to have little influence on the settlement rate as indicated
by the low rates for the two largest embankments, Eucembene and San Luis (main
Chapter 7: Deformation behaviour of embankment dams Page 7.130

section) dams, compared to some of the other case studies of much lower embankment
height.
For these embankments with very broad earthfill zones the development of the
phreatic surface and the influence of reservoir fluctuation on piezometric levels within
the embankment are considered to influence the post construction deformation
behaviour. For most of the case studies the earthfill zones are dominantly clayey and of
low permeability, and an important consideration for the deformation behaviour of these
embankments is the effect of the gradual development of the phreatic surface and
associated changes in effective stress conditions within the embankment. For a number
of the case studies (of dominantly clayey earthfills) piezometric records indicate the
phreatic surface took tens of years to reach steady levels in the central core region of the
broad earthfill core, whilst for more permeable earthfills steady levels were reached
within several years. Some examples of this are:
• At Belle Fourche dam (35 m high), constructed of medium plasticity clays placed on
the dry side of Standard optimum, borehole “well” records indicate it took more
than 10 to 15 years to establish steady levels in the upstream earthfill zone and more
than 20 years in the central embankment region.
• At Fresno dam (24 m high), constructed of low plasticity sandy clays placed at an
average of 0.2% dry of Standard optimum, pore water pressures reached close to
equilibrium conditions more than 4 years but less than 15 years after end of
construction (Daehn 1955).
• At Steinaker dam (49 m high), central earthfill zone of medium plasticity clays
placed on the dry side of Standard optimum, piezometer records indicate it took 15
to 30 years to establish steady levels in the central embankment region.
• At San Luis dam (115 m high), central earthfill zone of low to medium plasticity
clays placed on the dry side of Standard optimum, piezometer records indicate it
took 15 to 30 years to establish steady levels in the central embankment region.
• At the Horsetooth Reservoir embankments (Horsetooth, Dixon Canyon, Spring
Canyon and Soldier Canyon dams of 48 to 74 m height), constructed of
predominantly sandy clays placed on the dry side of Standard optimum, USBR
(1997) referred to the piezometer levels still rising more than 50 years after end of
construction.
• At Khancoban dam, earthfill zone of silty sands, reasonably steady levels were
established within 2 to 5 years after end of construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.131

For those embankments with cores of low permeability, reservoir fluctuation would
not be expected to greatly influence the crest settlement rate due to the likely limited
effect it would have on pore water pressures in the central region of the embankment.
However, for more permeable earthfills the reservoir operation may have a more
substantial influence on the post construction crest deformation behaviour. At Enders
dams the increased rate in settlement 8 to 10 years after the end of construction (Figure
7.60b) is coincident with the increased magnitude of fluctuation of the reservoir. In the
first 8 to 10 years of operation annual fluctuations were limited to less than about 3 m,
but after this period increased to 5 to 8 m (Figure 7.66). Piezometric records indicate
that pore water pressures within the central embankment region also fluctuated with
reservoir level but at reduced amplitude.
In comparison, the post construction settlement at the main section at San Luis dam
(Figure 7.60b) showed no significant change in rate of settlement beyond about 8 years
when the seasonal reservoir fluctuation increased substantially (Figure 7.67).
Piezometers in the central region of the main earthfill zone (of mostly low to medium
plasticity sandy clays) showed a negligible to very limited response to drawdown.

100 950

80
940
Settlement (mm) .

Crest - 4m upstream of axis (Stn 48.4)

Reservoir Level (m)


Crest - 4m dowmstream of axis (Stn 46)
60 20m downstream (Stn 44.5)
Reservoir 930

40

920
20

910
0

-20 900
0 5 10 15 20 25 30 35 40 45 50

Time (years since end of construction)

Figure 7.66: Post construction settlement and reservoir operation at Enders dam (data
courtesy of USBR).
Chapter 7: Deformation behaviour of embankment dams Page 7.132

500 166

141
400
Settlement (mm) .

Reservoir Level (m) .


116
300

91

200
66

Crest - downstream edge (Stn 110)


100 13m upstream (Stn 112)
41
68m downstream (Stn 112)
Reservoir
0 16
0 5 10 15 20 25 30 35

Time (years since end of construction)


Figure 7.67: Post construction settlement at main section and reservoir operation at San
Luis dam (data courtesy of USBR).

7.6.4 POST CONSTRUCTION HORIZONTAL DISPLACEMENT OF THE CREST NORMAL TO


THE DAM AXIS.

The post construction horizontal displacements of the crest are presented in Figure 7.68
to Figure 7.76 in the form of total displacement in millimetres versus time (in years
since end of construction) on a log scale. The case studies have been sorted based on
core width, core material type and placement moisture content into the following
figures:
• Figure 7.68 – thin to medium core widths, dry placed clay cores;
• Figure 7.69 – thin to medium core widths, wet placed clay cores;
• Figure 7.70 – thin to medium core widths, dry placed clayey sand to clayey gravel
cores;
• Figure 7.71 – thin to medium core widths, wet placed clayey sand to clayey gravel
cores;
• Figure 7.72 – thin to medium core widths, silty sand to silty gravel cores both dry
and wet placed;
• Figure 7.73 – clay cores of thick core width;
• Figure 7.74 – silty and clayey sand and gravel cores of thick core width;
• Figure 7.75 – very broad cores where the foundation has limited influence on
deformation; and
Chapter 7: Deformation behaviour of embankment dams Page 7.133

• Figure 7.76 – very broad cores where the foundation has or potentially has a
significant influence on settlement.

For most case studies the displacement includes the period of first filling, or at least
a substantial portion of it. For several embankments though first filling had been (or
was virtually) completed prior to the start of monitoring and includes, but is not limited
to, Nillahcootie, Eildon, Belle Fourche and Peter Faust dams.
For several of the embankments with very broad earthfill cores the distinction of
foundation influence could not be readily evaluated, as no data on the foundation
deformation was available. These cases include Pueblo, Dixon Canyon and Spring
Canyon dams.
The size of each figure has been set to the same margins for each plot area and the
time interval standardised to cover from 0.1 to 100 years to allow for visual comparison
between the figures. In most cases a standard vertical axis of 100 mm upstream (-100
mm) to 300 mm downstream (+300 mm) displacement has been used. For several plots
where larger displacements were recorded a broader vertical axis has been adopted.
Other points to note about the figures are explained in the preamble for the crest
settlement versus time plots (Section 7.6.3).

300
1
End of first filling

250
Displacement, mm (downstream +ive) .

200

150

100
5
5

50
4
3
2
0
4 6
7 7
6 2
-50

-100
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Bellfield (704) 2. Split Yard Creek (21) 3. Nillacootie (701) 4. Peter Faust (15)
5. Windemere (31) 6. Glenbawn Saddle (19) 7. Bjelke Petersen (13)

Figure 7.68: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.134

1000

End of first filling 1


Displacement, mm (downstream +ive) .

800

600
1

400
2

200 4
3
5
6 7
3
6
0

2
-200
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Djatilihur 2. El Infiernillo (M10) 3. Wivenhoe (SS26) 4. Talbingo (CS6)
5. Blowering (IVM A) 6. William Hovell (Ch. 503) 7. La Agnostura (Pnt. d)

Figure 7.69: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey earthfills.

300
1. Copeton (74)
2. Googong (S6)
250 3. Corin (24)
4. Wyangala (32)
Displacement, mm (downstream +ive) .

5. Fukada
200

End of first filling


150

3
1
100
2 1
3
50
4
5
0

4
-50

-100
0.1 1.0 10.0 100.0
Time (years since end of construction)
Figure 7.70: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of dry placed clayey sand and clayey gravel (SC/GC)
earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.135

500

End of first filling 9


9 7
400
Displacement, mm (downstream +ive) .

300
6
7
6
3
200
1
5
2
100

8 5 4
0

3
8 2
4
-100

-200
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Gepatsch (3) 2. Shimokotori 3. Seto 4. Dartmouth (CS7) 5. Thomson (CS6)
6. Kurokawa 7. Kisenyama 8. Rosshaupten 9. Beliche (J54)

Figure 7.71: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of wet placed clayey sand and clayey gravel (SC/GC)
earthfills.

1200
3

End of first filling


1000
Displacement, mm (downstream +ive) .

800

600
1
8
8
400

2 7 2
9
200
5
5
6
4
0

6, 4
-200
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Round Butte (19) 2. Ajuare 3. Svartevann (main section) 4. Parangana (Ch. 400) 5. Geehi (5)
6. Rowallan (Ch. 1500) 7. Mud Mountain 8. Cougar (TD-11) 9. Vatnedalsvatn

Figure 7.72: Crest displacement versus time for zoned embankments with thin to
medium width central core zones of silty sand and silty gravel (SM/GM) earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.136

350
9
9
300
End of first filling
Displacement, mm (downstream +ive) .

250
7
5
200
3 7

150
5 3

100

1
50
10 1

10
0
4
4 6 8
-50 6

2
2
-100
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Waipapa (DM16) 2. South Holston 3. Upper Yarra (M5) 4. Maroon (25) 5. Eppalock (CS2)
6. Burrendong (73) 7. Navajo (Stn 38) 8. Eildon (CS7) 9. Glenbawn (G2H) 10. San Justo (Stn 3.3)

Figure 7.73: Crest displacement versus time for zoned embankments with central core
zones of clayey earthfills of thick width (1 to 2.5H to 1V combined width).

300

End of first filling


250
Displacement, mm (downstream +ive) .

200
6

150

6
100
2
4 3 2
8 3
50
4

0
8
7
5
-50 1
7
1

-100
0.1 1.0 10.0 100.0
Time (years since end of construction)

1. Cardinia (SS13) 2. Greenvale (CS5) 3. Bradbury (Stn 19) 4. Benmore (DM16) 5. Tooma (IVM A)
6. Cherry Valley (Stn 20) 7. Mammoth Pool (B11) 8. Meeks Cabin (Stn 16.5)

Figure 7.74: Crest displacement versus time for zoned embankments with central core
zones of silty to clayey sand and gravel earthfills of thick width.
Chapter 7: Deformation behaviour of embankment dams Page 7.137

250

200
1
6 6
Displacement, mm (downstream +ive) .

150

100
9 3, 4 5
3 4
11 10
50
11
0 8

-50
7

-100

-150
2
End of first filling 7
-200
9 2

-250
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Belle Fourche (MP10) 2. Mita Hills (A4) 3. Pueblo, right emb. (Stn 46) 4. Pueblo, left emb. (Stn 77.5)
5. Tullaroop (CS2) 6. San Luis (Stn 110) 7. San Luis, slide area (Stn 138) 8. Cairn Curran (C3)
9. Rector Creek (Stn 16) 10. Eucembene (15) 11. Khancoban (12/13)

Figure 7.75: Crest displacement versus time for embankments with very broad core
zones (> 2.5H to 1V width) and with limited foundation influence.

300

End of first filling 1 to 4 5


1
250
Displacement, mm (downstream +ive) .

200

150
5

100

6 2

50

6
0
4
3
-50

-100
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Horsetooth (Stn 10) 2. Dixon Canyon (Stn 9) 3. Soldier Canyon (Stn 11) 4. Spring Canyon (Stn 5)
5. San Luis (Stn 86) 6. Steinaker (Stn 15)

Figure 7.76: Crest displacement versus time for embankments with very broad core
zones (> 2.5H to 1V width) and with potentially significant foundation influence on the
embankment deformation.
Chapter 7: Deformation behaviour of embankment dams Page 7.138

A typical pattern of crest displacement behaviour is difficult to define, however, for a


large number of the case studies:
• A large portion of the crest displacement occurs during first filling, in most cases the
displacement being downstream.
• Post first filling, most cases show a slight to steady downstream trend, approaching
very small to near zero displacement rates in the long-term (after 10 to 20 years).
• Fluctuations about the general trend occur due to reservoir fluctuation, with
upstream displacement on drawdown and downstream displacement on first filling.

Plots of crest displacement versus reservoir level presented for a number of


embankments in the literature highlight this general trend.

(a) Displacement during first filling


The lateral displacement at end of first filling was discussed in Section 7.5.3. Plots of
the displacement at end of first filling versus embankment height were presented for the
crest (Figure 7.38) as well as the upstream and downstream slopes. Typical ranges of
the crest displacement at end of first filling are given in Table 7.17 sorted on core width,
core properties (material type and moisture content at placement) and the material type
and placement method of the downstream shoulder.
For most cases the crest displacement is dominantly downstream reaching a
maximum at the end of first filling.
As discussed in Section 7.5.2, central core earth and rockfill embankments with wet
placed cores of low undrained strength and upstream rockfill susceptible to collapse
settlement on saturation had the potential to deform upstream during the early stages of
reservoir filling followed by downstream displacement as the reservoir approached full
supply level. Nobari and Duncan (1972b) proposed this concept to explain the crest
displacement behaviour at El Infiernillo dam. From the crest displacement versus time
plots it is apparent that this behaviour also occurred at Gepatsch and Thomson dams,
and possibly also at Blowering and Dartmouth dams; all of which are central core earth
and rockfill embankments with thin to medium width cores of wet placed clays, clayey
sands or clayey gravels.
Chapter 7: Deformation behaviour of embankment dams Page 7.139

(b) Displacement Post First Filling


Post first filling the general trend for crest displacement, as indicated in the figures, is
for a steady or gradual downstream displacement with time approaching low or near
zero displacement rates long-term.
Figure 7.77 presents the lateral displacement post first filling versus embankment
height. The displacement data is for at least five years after first filling (up to 30 years
or more for some cases). Where possible the long-term data point has been taken for
the reservoir at close to full supply level. For Belle Fourche dam the data is only for
monitoring from the period 75 to 85 years post construction from the closure section,
and the actual post first filling displacement is likely to be much greater than shown for
this region of the embankment. For Matahina dam the post first filling crest
displacement is prior to the earthquake in 1987.
The data shows that for most cases the displacement post first filling is generally in
the range 35 mm upstream (-35 mm) to 100 to 150 mm downstream, indicating that post
first filling crest displacements are generally small and are not dependent on dam
height. Displacements outside this range may be indicative of “abnormal” deformation
behaviour. Several of these case studies are discussed further in Sections 7.10 and 7.11.

450

Djatilihur El Infiernillo
Displacement post first filling (mm) .

Rector Creek
(downstream is positive) .

300
Matahina
Kisenyama
Eppalock
San Luis - Stn 86

150 Belle Fourche


(75 to 85 yrs
period only)

Blowering

-150
0 50 100 150 200 250
Dam Height (m)
Thin to thick clay cores, fluctuating reservoir Thin to thick SM/GM cores, fluctuating reservoir
Thin to thick clay cores, slow or steady reservoir Thin to thick SM/GM cores, slow or steady reservoir
Thin to thick SC/GC cores, fluctuating reservoir Very broad cores
Thin to thick SC/GC cores, slow or steady reservoir Very broad cores, possible foundation influence

Figure 7.77: Crest displacement normal to dam axis post first filling
Chapter 7: Deformation behaviour of embankment dams Page 7.140

7.6.5 POST CONSTRUCTION DEFORMATION OF THE MID TO UPPER DOWNSTREAM


SLOPE

Post construction deformation data for the mid to upper region of the downstream slope
has been presented in previous sections, including:
• Lateral displacement normal to the dam axis at the end of first filling in Section
7.5.3. Figure 7.39 presents the lateral displacement at end of first filling versus
embankment height and Figure 7.42 versus the downstream slope angle.
• Total settlement plots at 3 years (Figure 7.49) and 10 years (Figure 7.50) after the
end of embankment construction in Section 7.6.2. Table 7.20 summarises the
typical range of settlement sorted based on material type and placement method of
the downstream shoulder material. The data was plotted against the height from the
SMP to foundation level.

Additional data on the post construction deformation of the mid to upper


downstream slope presented below includes:
• Figure 7.78 – Long-term settlement rate (% per log cycle of time) versus the height
from the SMP to foundation level;
• Table 7.22, summarising the long-term settlement rate data;
• Figure 7.79 – Settlement, as a percentage of the height from the SMP to foundation
level, versus log time for selected case studies; and
• Figure 7.80 to Figure 7.82 – Lateral displacement versus log time for selected case
studies.

A full complime nt of the total settlement (at 3, 10 and 20 to 25 years after


construction), settlement versus time and displacement versus time plots for the mid to
upper region of the downstream shoulder are presented in Section 2.1 of Appendix F.
For all plots the data has been sorted based on material type forming the
downstream shoulder; rockfill, gravels or earthfill. Embankments with rockfill in the
downstream shoulder have been further sorted based on compaction rating. For the total
deformation and settlement rate plots, case studies where the rockfill was placed
without the addition of water have been highlighted.
Chapter 7: Deformation behaviour of embankment dams Page 7.141

Table 7.22: Range of long-term settlement rate of the downstream shoulder.


No. Long-term Settlement
Material Compaction
Cases Rate *1, *2 Comments
Type Rating *3
(% per log cycle of time)
well 10 0.0 to 0.33
reas to well 9 0.04 to 0.31 (most = 0.15)
reas 4 0.10 and 0.31 Ataturk = 0.96%,
Rockfill
Gepatsch = 0.99%
poor 7 0.10 to 0.25
poor – dry 6 0.20 to 0.75
Gravels - 6 0.02 to 0.065 0.24% at Meeks Cabin
Earthfills - 19 0.0 to 0.40 Refer Figure 7.78 for
outliers.
Note: *1 Excludes possible outliers.
*2 Settlement as a percentage of the height from the SMP to foundation level.
*3 compaction rating of rockfill; well = well-compacted, reas to well = reasonably to well
compacted, reas = reasonable compaction, poor = poorly compacted, poor – dry = dry placed
and poorly compacted. Refer Section 1.3.3 for definitions of the terms.

Most of the essential features of the post construction deformation behaviour of the
downstream slope are captured in the plots and tables of total deformation and long-
term settlement rate.
The typical trend for post construction settlement is for near constant or slightly
increasing rate of settlement with the log of time. Periods of higher settlement rate are
observed for some of the embankments with poorly and reasonably compacted rockfill
in the downstream shoulder. Notable features of the post construction settlement are:
• In terms of total settlement:
− Settlements in the order of 1 to 2% are observed for poorly compacted rockfills.
Greater settlements are observed for the dry placed, poorly compacted rockfills.
− For reasonably compacted rockfills the range of settlement is quite broad, from
0.1% up to 1.0%, but the number of cases is limited. Settlements toward the
upper range are observed for dry placed and/or weathered rockfills, where
settlements due to collapse compression are likely to be significant.
− Much lower settlements, generally less than 0.5 to 0.7% at ten years after
construction, are observed for well and reasonably to well compacted rockfills,
and compacted earthfills.
Chapter 7: Deformation behaviour of embankment dams Page 7.142

− Very low settlements (less than 0.25% at 10 years) are observed for
embankments with gravel shoulders.
• In terms of the long-term settlement rate:
− For most cases the long-term settlement rate is less than 0.4% per log cycle of
time.
− Higher settlement rates are observed for a number of earthfill embankments. At
Belle Fourche dam (within the closure section) the very high rate (close to 4%
per log cycle) is a clear outlier and indicative of “abnormal” behaviour.
− Higher settlement rates apply for poorly compacted rockfills, ranging from 0.2%
to 0.75% per log cycle of time, with the dry placed cases generally at the higher
end of the range.
− Higher settlement rates also apply for reasonably compacted rockfills that are
susceptible to significant settlements due to collapse compression from wetting
(dry placed and/or weathered rockfills). For both Ataturk and Gepatsch dams
the indicated rates may be an over-estimate because they have been derived from
data inclusive of the first few years after construction.
− Very low settlement rates apply for embankments with gravel downstream
shoulders (< 0.10% per log cycle of time).
− Reservoir operation post first filling has a limited effect on the settlement rate.

As indicated, the post construction settlement behaviour of the downstream slope


within the closure section at Belle Fourche dam is a clear outlier and indicative of
“abnormal” behaviour. Other case studies where the magnitude of settlement or long-
term settlement rate is possibly indicative of “abnormal” behaviour includes
Svartevann, Horsetooth, Dixon Canyon, Spring Canyon, Srinagarind, Pueblo (both the
left and right embankments), Ataturk and possibly Gepatsch dams. A number of these
cases are discussed further in Sections 7.10 and 7.11, and in Appendix G.
Chapter 7: Deformation behaviour of embankment dams Page 7.143

a) 5

Belle Fourche (closure section)


4
Long-term Settlement Rate .
(% per log time cycle) .

0
0 25 50 75 100 125 150 175
Height from SMP to foundation level (m)
Gravels, fluctuating reservoir gravels, steady reservoir
Earthfill, steady reservoir Earthfill, fluctuating reservoir
Rockfill, well-compacted - fluct. reservoir Rockfill, well-compacted - steady reservoir
Rockfill, reasonably to well compacted - fluct. reservoir Rockfill, reasonably to well compacted - steady reservoir
Rockfill, reasonably compacted - fluct. reservoir Rockfill, reasonably compacted - steady reservoir
Rockfill, poorly compacted - fluct. reservoir Rockfill, poorly compacted - steady reservoir

b) 1.2

Dixon Canyon

1
Gepatsch Ataturk
Long-term Settlement Rate .

Horsetooth Spring Canyon


(% per log time cycle) .

0.8 Pueblo (left


abut.)
Eppalock Pueblo (right abut.)
Upper Yarra
0.6
Bellfield

0.4

0.2

indicates rockfill placed


without the addition of water
0
0 25 50 75 100 125 150 175
Height from SMP to foundation level (m)
Note: refer to (a) above for legend.

Figure 7.78: Long-term settlement rates for the downstream slope (mid to upper region)
versus embankment height; (a) all data, and (b) data excluding Belle Fourche.
Chapter 7: Deformation behaviour of embankment dams Page 7.144

0.0%
10
11
Settlement (% of height from SMP to foundation) .
8
12
6
1 7
0.5%
5 influence of earthquake
9 1 at Matahina dam
4
End of first filling

1.0%
3

2
2
9
1.5%
Downstream Shoulder:
Cases 1 to 4 - poorly compacted rockfill
Cases 5 and 6 - well-compacted rockfill
Cases 7 and 8 - reasonably to well-compacted rockfill
Case 9 - reasonably compacted rockfill
2.0% Case 10 - gravels
Cases 11 & 12 - earthfill (Case 12 very broad earthfill core)
3

0.1 1.0 10.0 100.0

Time (years since end of construction)

1. El Infiernillo (M23) 2. Beliche (J40) 3. Eildon (SS4) 4. Matahina (21) 5. Dartmouth (22) 6. Talbingo (33)
7. Copeton (101) 8. Geehi (18) 9. Gepatch (4) 10. Trinity (Stn 13) 11. Navajo (Stn 41) 12. San Luis (Stn 112)

Figure 7.79: Post construction settlement of the downstream shoulder (mid to upper
region) for selected case studies.

The displacement versus log time for selected case studies are presented in Figure 7.80
to Figure 7.82, and for all case studies in Figures F2.12 to F2.19 in Section 2.1 of
Appendix F. As for the crest displacement, a typical pattern of behaviour for the
horizontal displacement of the downstream slope is difficult to define. The general
trend is for downstream displacement on first filling and continued downstream
displacement long-term, although for several case studies the displacement is slightly
upstream. Several trends that are evident are:

(a) Earthfill and zoned embankments with very broad core widths
For embankments of very broad core width the displacement (Figure 7.80) shows a
steady downstream rate of displacement with log time, the rate long-term usually being
higher than in the early years after construction. First filling generally has little
influence on the deformation behaviour, except at Horsetooth dam where the
displacement was influenced by the orientation of the cut-off trench and the large
deformation of the foundation.
Chapter 7: Deformation behaviour of embankment dams Page 7.145

The limited of influence of first filling is largely due to the location and orientation
of the applied water load on the upstream face of the very broad earthfill zone. The
gradual development of the phreatic surface within the embankment over many years
(refer Section 7.6.3, point f) is considered a factor in the observation of higher rates of
displacement post first filling (rate per log time).
The magnitude of post construction displacement of the downstream slope is small
for earthfill embankments and embankments with very broad core widths.
Displacements at 25 to 45 years after construction range from 10 mm to 200 mm
downstream, or 0.05% to 0.30% of the embankment height (excludes Horsetooth and
Dixon Canyon dams). The displacement at Dixon Canyon dam is much larger at almost
300 mm (0.46% of the embankment height), and at Horsetooth dam is 260 mm, 180 mm
of this occurring on first filling due mainly to foundation influence. At Belle Fourche
dam the very high rate of displacement of the downstream slope in the closure section
(Figure 7.80), where 90 mm was measured over the period 75 to 85 years after
construction, is considered to be “abnormally” high.

300
2
1. Pueblo, right emb. (Stn 48.5)
2. Dixon Canyon (Stn 10)
250 3. Belle Fourche (MP16)
Displacement, mm (downstream +ive) .

4. Tullaroop (SS2)
5. San Luis (Stn 112)
6. Soldier Canyon (Stn 11)
200 7. Eucembene (9)
8. Khancoban (29) 5

150
End of first filling
1

2 6
100
3
4

50
1 7

8
0

6
-50
0.1 1.0 10.0 100.0
Time (years since end of construction)
Figure 7.80: Post construction displacement of the downstream shoulder (mid to upper
region) for selected case studies of embankments with very broad core widths.
Chapter 7: Deformation behaviour of embankment dams Page 7.146

(b) Central core earth and rockfill embankments


For central core earth and rockfill embankments the general trend of displacement of the
downstream slope is for higher rates of downstream displacement (rate per log time)
during the first few years after end of construction decreasing to much lower rates long-
term (Figure 7.81). Two explanations for this are; firstly, for thin to medium central
cores the water load has a high horizontal component and is applied close to the centre
of the embankment resulting in an increase in lateral stress in the downstream shoulder
on first filling and downstream displacement of the downstream shoulder. Secondly,
wetting of the downstream rockfill from rainfall infiltration or tail water inundation
causing large deformations in those rockfills that are susceptible to large deformations
due to collapse compression. Rockfills most susceptible, as previously identified, are
those that are dry placed and reasonably to poorly compacted, or where the rock
substance strength is susceptible to large loss of strength on wetting.
The magnitudes of displacement of the downstream shoulder is generally in the
range:
• For well and reasonably to well compacted rockfills, displacements long-term (more
than 10 to 20 years after construction) are typically less than 0.20% of the
embankment height, or less than about 100 to 300 mm. Larger displacements, up to
0.25% to 0.40% of the embankment height, can occur where the rockfill has been
placed dry, the rock substance strength is susceptible to large strength loss on
wetting and/or the outer zone of the rockfill has been placed in thicker layers (i.e.
reasonably compacted).
• For reasonably and poorly compacted rockfills, displacements can reach values up
to 1.0% to 1.6% of the embankment height long-term. Those most susceptible to
the larger displacements are rockfills that have been dry placed and poorly
compacted. Post construction displacements close to and in excess of 1 m have
been measured at Ataturk (184 m high), Eildon (80 m high), Gepatsch (153 m high)
and Svartevann (129 m high) dams. As a percentage of the embankment height the
highest displacements have been measured at the 80 m high Eildon dam (1.52% of
the embankment height) and the 46 m high Eppalock dam (1.63% of the
embankment height). For both these dams the rockfill was dry placed and poorly
compacted.
Chapter 7: Deformation behaviour of embankment dams Page 7.147

1250
Downstream Shoulder: 3
Cases 1 to 3 - poorly compacted rockfill
Cases 4 to 6 - well-compacted rockfill
Cases 7 to 9 - reasonably to well-compacted rockfill
Displacement, mm (downstream +ive) .

1000 Cases 10 and 11 - reasonably compacted rockfill

11 11 10
End of first filling

750

3 10
500
7
1 7
2

9 4
2 5
4
250
1

8 8
6 9
0
6
0.1 1.0 9 10.0 100.0
Time (years since end of construction)
1. El Infiernillo (M23) 2. Beliche (J40) 3. Eildon (SS4) 4. Dartmouth (22) 5. Talbingo (33) 6. Wivenhoe (34)
7. Copeton (101) 8. Geehi (18) 9. William Hovell (Ch. 503) 10. Gepatch (4) 11. Svartevann (EL 870 m)

Figure 7.81: Post construction displacement of the downstream shoulder (mid to upper
region) for selected case studies of central core earth and rockfill embankments.

(c) Zoned embankments with compacted earthfill and gravels in the downstream
shoulder
For zoned embankments with compacted earthfills and gravels in the downs tream
shoulder, the post construction displacement (Figure 7.82) is relatively small at less than
about 150 to 250 mm (less than 0.15 to 0.20% of the embankment height) some 20 to 30
years after construction. For most of the case studies the displacement rate (per log
cycle of time) decreases with time approaching near zero or very low values long-term.
In general, first filling has a limited to negligible influence on the displacement, except
at La Angostura dam, where the embankment comprised a thin clay core, and possibly
Trinity dam.
Chapter 7: Deformation behaviour of embankment dams Page 7.148

300
Cases 1 to 6 - earthfills in downstream shoulder
Cases 7 to 11 - gravels in the downstream shoulder

250
Displacement, mm (downstream +ive) .

8
End of first filling
8
200

150
2
7
2 1 1
100
3
10 4
50
3

5
6
0

11 10 9
5 4 11
9
-50
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Navajo (Stn 41) 2. Cardinia (26) 3. Greenvale (SS16) 4. Bradbury (Stn 18) 5. San Justo (Stn 3.0) 6. Tooma (7)
7. La Agnostura (Pnt. f) 8. Trinity (Stn 13) 9. Meeks Cabin (Stn 14) 10. Benmore (DM16) 11. Ruataniwha (DM5)

Figure 7.82: Post construction displacement of the downstream shoulder (mid to upper
region) for embankments with compacted earthfills and gravels in the downstream
shoulder.

7.6.6 POST CONSTRUCTION DEFORMATION OF THE UPPER UPSTREAM SLOPE AND


UPSTREAM CREST

Post construction deformation data of the upper upstream slope and upstream crest
region has been presented in previous sections, including:
• Lateral displacement normal to the dam axis at the end of first filling in Section
7.5.3. Figure 7.40 presents the lateral displacement at end of first filling versus
embankment height.
• Total settlement plots at 3 years (Figure 7.51) and 10 years (Figure 7.52) after the
end of embankment construction in Section 7.6.2. Table 7.20 summarises the
typical range of settlement sorted based on material type and placement method of
the upstream shoulder material. The data was plotted against the height from the
SMP to foundation level.
Chapter 7: Deformation behaviour of embankment dams Page 7.149

Additional data on the post construction deformation of the upper upstream slope
and upstream crest region presented below includes:
• Figure 7.83 – Long-term settlement rate (% per log cycle of time) versus
embankment height. The calculated rate excludes short periods of increased
settlement rate post first filling shown for some of the embankments, such as on
drawdown at 17 years for Djatiluhur dam (Figure 7.84);
• Table 7.23, summarising the long-term settlement rate data;
• Figure 7.84 and Figure 7.85 – Settlement (% of height from SMP to foundation)
versus log time for poorly compacted rockfills and selected other case studies
respectively; and
• Figure 7.86 and Figure 7.87 – Lateral displacement versus log time for selected case
studies.

A full compliment of the total settlement (at 3, 10 and 20 to 25 years after


construction), settlement versus time and displacement versus time plots are presented
in Section 2.2 of Appendix F.
For all plots the data has been sorted based on material type forming the upstream
shoulder; rockfill, gravels or earthfill. Embankments with rockfill in the downstream
shoulder have been further sorted based on compaction rating, and for the total
deformation and settlement rate plots case studies where the rockfill was placed without
the addition of water have been highlighted.
Several trends are apparent for the post construction settlement for the upper
upstream shoulder and upstream crest region. For most case studies the usual trend is
for near constant or slightly increasing rate of settlement with the log of time. In some
cases a decrease or increase in settlement rate is observed post first filling or the
settlement versus time curve shows a slight curvature of increasing rate with log time,
but the transition is generally smooth. This type of settlement time is generally
observed for:
• Zoned embankments with permeable earthfills or gravels in the upstream shoulder.
• Embankments with very broad earthfill cores. Horsetooth is an exception, but this
may be due to foundation influence.
• Well-compacted and reasonably to well compacted rockfills in the upstream
shoulder. Exceptions are La Angostura (possibly affected by the outer dumped
rockfill zone), Dartmouth, Blowering, Wyangala and La Grande 2 dams.
Chapter 7: Deformation behaviour of embankment dams Page 7.150

• A limited number of case studies with reasonably and poorly compacted rockfills in
the upstream shoulder, e.g., Tooma and Burrendong dams, both of which comprised
dumped and sluiced rockfills placed in 1.8 to 3 m lifts.

For a number of the poorly and reasonably compacted rockfills higher settlement
rates or periods of higher settlement rate during first filling followed by a relatively
steady, reduced settlement rate post first filling are observed. Relatively large collapse
settlement of the upstream rockfill on first filling is likely to be the cause of this
settlement trend.
A limited number of the case studies show periods of higher settlement rate post
first filling, generally occurring on drawdown. These cases include Dartmouth,
Blowering, Wyangala, Cougar, Djatiluhur, Beliche, Eildon and Eppalock dams, most of
which are represented in Figure 7.84 or Figure 7.85. This type of settlement trend is
potentially indicative of “abnormal” deformation behaviour. Assuming the rockfill is of
high permeability, close to the highest vertical stress levels are reached in the wetted
rockfill during first filling. On drawdown the vertical stresses in the upstream rockfill
will increase, but it is essentially in re-loading. Effective stress levels greater than
previously experienced by the wetted rockfill can potentially occur in the lateral
direction, normal to the dam axis, as the water load acting on the core is reduced and
therefore transferred onto the upstream rockfill. Another possible mechanism is
localised instability of the core on drawdown due to the reduction in lateral stresses
acting on the core on drawdown. These mechanisms, along with a number of the cases
studies, are discussed further in Section 7.10.
Table 7.20, Figure 7.51 and Figure 7.52 provide information on the general range of
settlement of the upper upstream slope and upstream crest region sorted based on
material type and placement method of the upstream shoulder material. In summary,
the general ranges of settlement as a percentage of the height from the SMP to
foundation level are:
• In the order of 1 to 2% for poorly compacted rockfills. Greater settlements are
observed for the dry placed, poorly compacted rockfills. As shown in Figure 7.84 a
large portion of the settlement occurs on first filling as collapse type settlement.
• For several embankments with reasonably compacted rockfill in the upstream
shoulder, notably Gepatsch and Cougar dams, greater settlements are observed and
Chapter 7: Deformation behaviour of embankment dams Page 7.151

are probably due to large collapse type settlements in the dry placed rockfill at
Gepatsch dam and the weathered rockfill at Cougar dam.
• Less than 0.5% to 0.7% settlement at 10 years for the well and reasonably well
compacted rockfills, and for embankments with earthfill zones in the upstream
shoulder, including earthfill embankments.
• Less than 0.25% settlement at 10 years for embankments with gravel shoulders, but
there are only three cases with deformation data from the database.

The long-term settlement rate data (Figure 7.83 and Table 7.23) indicates:
• For most cases the long-term settlement rate is less than 0.4 % per log cycle of time.
• Higher rates, up to 0.8% per log cycle of time, are observed for a number of
embankments. There is no apparent consistency in material type and reservoir
operation for these cases.
• Embankment height, reservoir operation, material type and compaction rating do not
appear to have a significant influence. Gravels possibly have a lower settlement
rate, but there are only two cases.

Table 7.23: Range of long-term settlement rate of the upper upstream slope and
upstream crest region.
No. Long-term Settlement
Material Compaction
Cases Rate *1 *2 Comments
Type Rating *3
(% per log cycle of time)
well 8 0.05 to 0.70 (most < 0.50) Dartmouth = 0.67%,
Glenbawn Saddle = 0.675%
Rockfill reas to well 9 0.10 to 0.56 (most < 0.50)
reas 3 < 0.55
poor 7 0.10 to 0.82 (most < 0.60) Djatiluhur = 0.82%
Gravels - 2 < 0.21
Earthfills - 18 0.1 to 0.60 Belle Fourche (closure
section) = 2%,
Dixon Canyon = 1.38%
Note: *1 Excludes possible outliers.
*2 Settlement as a percentage of the height from the SMP to foundation level.
*3 compaction rating of rockfill; well = well-compacted, reas to well = reasonably to well
compacted, reas = reasonable compaction, poor = poorly compacted, poor – dry = dry placed
and poorly compacted. Refer Section 1.3.3 for definitions of the terms.
Chapter 7: Deformation behaviour of embankment dams Page 7.152

Potential outliers in terms of magnitude of settlement of the upstream shoulder include


Dixon Canyon, Spring Canyon and Horsetooth dams. Belle Fourche dam (within the
closure section) is a clear outlier in terms of long-term settlement rate as is probably
Dixon Canyon dam. The settlement rates are on the high side at Glenbawn Saddle and
Dartmouth dams compared to other well-compacted rockfills. Several of these cases are
discussed further in Section 7.10 and 7.11.

2.5

Belle Fourche (closure section)


2
Long-term Settlement Rate .
(% per log time cycle) .

1.5
Dixon Canyon

0.5

0
0 25 50 75 100 125 150 175
Height from SMP to foundation level (m)

Djatilihur
0.8
Long-term Settlement Rate .
(% per log time cycle) .

Glenbawn Saddle Soldier Canyon Dartmouth

Spring Canyon Upper Yarra


0.6

William Hovell

0.4

0.2

0
0 25 50 75 100 125 150 175
Height from SMP to foundation level (m)
Gravels, fluctuating reservoir Rockfill, well-compacted - steady reservoir
Rockfill, well-compacted - fluct. reservoir Rockfill, reasonably to well compacted - steady reservoir
Rockfill, reasonably to well compacted - fluct. reservoir Rockfill, poorly compacted - steady reservoir
Rockfill, reasonable compaction - fluct. reservoir Earthfill, steady reservoir
Rockfill, poorly compacted - fluct. reservoir Earthfill, fluctuating reservoir

Figure 7.83: Long-term settlement rates of the upper upstream slope to upstream crest
region of earthfill and earth-rockfill embankments.
Chapter 7: Deformation behaviour of embankment dams Page 7.153

0.00%
8
Settlement (% height from SMP to foundation) .
16
0.25%
6
8
4 11
0.50%
1 3 2 3 6

0.75%

1.00%
9 13 15
12 14
10
2
1.25%

1.50%
9

7
1.75%
5
End of first filling
2.00%
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Djatilihur 2. Djatilihur (C9) 3. El Infiernillo (IVM D1) 4. Netzahualcoyotl (14-2u) 5. Beliche (M54) 6. Upper Yarra (1U)
7. Eildon (SS10) 8. Burrendong (103) 9. Eppalock (SS2) 10. Glenbawn (L3M) 11. Nillacootie (6U) 12. Mud Mountain
13. Cherry Valley 14. Nottely 15. Watuaga 16. Tooma (25)

Figure 7.84: Post construction settlement of the upper upstream slope to upstream crest
region for embankments with poorly compacted rockfill in the upstream slope.

0.0% 11
9
12
Settlement (% of height from SMP to foundation) .

0.2% 6
6
11

0.4% 9
2
5 10
10 3
0.6% 7
4

4 5
0.8%
2
End of first filling 1 1

1.0%

8
Upstream Shoulder:
1.2% Cases 1 to 4 - well compacted rockfill
Cases 5 and 6 - reasonably to well compacted rockfill
Cases 7 and 8 - reasonably compacted rockfill
1.4% Case 9 - gravels
Cases 10 to 12 - earthfill (Cases 11 &12 very broad
earthfill core)
12
1.6%
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Dartmouth (32) 2. Blowering (17) 3. Talbingo (14) 4. Glenbawn Saddle (12) 5. Copeton (54) 6. Geehi (32)
7. Cougar (TD-12) 8. Gepatsch (1) 9. Trinity (Stn 12) 10. Navajo (Stn 41) 11. San Luis (Stn 112) 12. Spring Canyon (Stn 5)

Figure 7.85: Post construction settlement of the upper upstream slope to upstream crest
region for selected case studies (excluding poorly compacted rockfills).
Chapter 7: Deformation behaviour of embankment dams Page 7.154

The horizontal displacement versus time of the upper upstream slope and upstream crest
region for selected case studies is presented in Figure 7.86 and Figure 7.87. The data
for all case studies is presented in Figures F2.30 to F2.36 in Section 2.2 of Appendix A.
As shown, there is a large variation in the magnitude and direction of the measured
displacements. The data presented below has been sub-divided into two groups, those
embankments with rockfills in the upstream shoulder, and those with earthfills,
including gravels, in the upstream shoulder.

(a) Horizontal displacement of embankments with rockfill upstream shoulders


Figure 7.86 presents the horizontal displacement of the upper upstream slope and
upstream crest region of selected case studies with rockfill in the upstream shoulder.
The case studies have been selected covering the range of rockfill compaction rating
from well compacted through to poorly compacted.
For rockfills that are well and reasonably to well compacted, displacements of the
upstream shoulder during and after first filling are generally relatively small. For some
case studies a large portion of the displacement occurs on first filling (e.g. for Talbingo
dam), whilst for others first filling has a limited influence. Post first filling
displacements are small in magnitude and long-term the displacement rate (per log cycle
of time) approaches a small to near zero value, with variations about the trend due to
reservoir fluctuation. The typical range of displacement for these case studies is:
• At more than 10 to 20 years after construction, displacements range from -0.10% to
0.15% of the embankment height (or -75 mm to 200 mm).
• Additional displacement post first filling ranges from -0.05% to 0.05% of the
embankment height (-50 to 50 mm).

For rockfills that are reasonably to poorly compacted, displacements can be much
larger reaching magnitudes of close to 1 metre, or close to 1% of the embankment
height. The case studies with large displacements tend to be dry placed and reasonably
to poorly compacted rockfills. A large portion of the displacement occurs on first filling
and long-term the displacement rate (per log cycle of time) approaches a small to near
zero value, with variations about the trend due to reservoir fluctuation. The typical
range of displacement for these case studies is:
• At more than 10 to 20 years after construction, displacements range from -0.70% to
0.80% of the embankment height (-600 mm to 1000 mm for the case studies). For
Chapter 7: Deformation behaviour of embankment dams Page 7.155

most of the case studies the long-term total displacement is less than about 0.2% to
0.3% of the embankment height, either upstream or downstream. Large
displacements were measured at Eppalock, Gepatsch, Svartevann and Glenbawn
dams, all dry placed rockfills.
• Displacements post first filling generally range from -0.15% to 0.05% of the
embankment height.

Factors affecting the magnitude of displacement of the upstream slope are likely to
include:
• The magnitude of differential deformation between the upstream shoulder and the
core or upstream transition zone of the embankment. For rockfills susceptible to
large deformations on collapse compression, differential deformations on first filling
can be large, and the vector of deformation of the upstream shoulder will be affected
by the internal zoning geometry of the embankment.
• Zoning of the rockfill in the upstream shoulder. Where placement methods, rock
types or the degree of weathering differs for different rockfill zones in the upstream
slope then differential deformations can develop between these rockfill zones and
therefore influence the magnitude and direction of the displacement.
• The deformation of the core and downstream shoulder of the embankment. The
closer the SMP on the upstream slope is located to the crest the greater the influence
the displacement of the core and downstream shoulder will be. In the case of
Svartevann and La Grande 2 dams, large downstream displacement of SMPs on the
crest (both the upstream and downstream edge) and downstream shoulder occurred
on first filling.
• The slope of the upstream shoulder will have some influence.

For a number of embankments the long-term trend of displacement shows several


trends that are potentially indicative of “abnormal” behaviour. These include,
continuing high or increasing rates of displacement (rate per log of time), and non-
recoverable displacements (usually upstream) on drawdown. Examples of this
behaviour include Blowering, Eppalock and possibly Eildon dams as shown in Figure
7.86. Other examples of this type of behaviour are observed at Copeton and Wyangala
dams. A number of these cases are discussed further in Section 7.10 and in Appendix
G.
Chapter 7: Deformation behaviour of embankment dams Page 7.156

750

6 6
End of first filling
600
Displacement, mm (downstream +ive) .

450
7

7 9 1
300

11 11
4 3
150 9
1

4
0
5

-150 5 12
10 2
10
12
-300
Upstream Shoulder:
Cases 1 to 3 - well compacted rockfill
Cases 4 to 6 - reasonably to well compacted rockfill 8 2
-450
Cases 7 and 8 - reasonably compacted rockfill
Cases 9 to 12 - poorly compacted rockfill 8
-600
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Dartmouth (32) 2. Blowering (17) 3. Talbingo (14) 4. Geehi (32) 5. Maroon (SS13) 6. La Grande 2 (23)
7. Cougar (TD-12) 8. Gepatsch (1) 9. Beliche (M54) 10. Eppalock (SS2) 11. Eildon (SS10) 12. Glenbawn (L3M)

Figure 7.86: Post construction lateral displacement of the upper upstream slope and
upstream crest region for selected case studies of embankments with rockfill in the
upstream shoulder.

(b) Horizontal displacement of embankments with earthfills and gravels in the


upstream shoulder
Embankments included in this sub-group include earthfill embankments, embankments
with very broad core width, and zoned embankments of thin to thick core width with
earthfills or gravels in the upstream shoulder. The displacement versus time for selected
case studies is shown in Figure 7.87.
For most case studies the magnitude of post construction displacement is relatively
small, and is either in an upstream direction or downstream direction. Long-term the
displacement rate (per log cycle of time) approaches very small (either up or
downstream) to near zero average values with fluctuation due to operation of the
reservoir. At more than 20 to 30 years after construction, the magnitude of
displacement is generally in the range -0.10% to 0.12% of the embankment height, or
from 100 mm upstream (-100 mm) to 200 mm downstream.
The influence of the compressible foundation is evident at Horsetooth dam where a
large downstream displacement occurred on first filling. The deformation of the
Chapter 7: Deformation behaviour of embankment dams Page 7.157

compressible foundation has possible also influenced the displacement of the section of
San Luis dam over the broad alluvial plain (San Luis – Stn 80 in Figure 7.87), as well as
Steinaker dam. For these two dams the displacements post first filling were larger than
at other dams.
A number of case studies show the initial displacement on and after first filling to be
upstream, and then long-term changes to downstream. These case studies include;
Navajo, Dixon Canyon, San Luis (main section and slide area), Spring Canyon, Soldier
Canyon and the right abutment embankment at Pueblo dam. This effect may be due to
development of the phreatic surface or possibly softening within the embankment.
The high rate of displacement of the upstream shoulder within the closure section at
Belle Fourche dam many years after construction is likely to be indicative of
“abnormal” behaviour. The deformation behaviour at Dixon Canyon dam, showing a
large upstream displacement on and after first filling followed by downstream
displacement, is also potentially “abnormal” when compared to that at similar
embankments.

300

End of first filling


9 7
200
Displacement, mm (downstream +ive) .

2
9
2
7
100
1
4
1
4
0
8
3

6
-100

3 5
10
6
-200
Note: Cases 1 to 3 are zoned embankments of thin to thick core 10
width with earthfills and gravels in the upstream shoulder.
Cases 4 to 10 are embankments with cores of very broad width.
-300
0.1 1.0 10.0 100.0
Time (years since end of construction)
1. Bradbury (Stn 19) 2. Trinity (Stn 12) 3. Navajo (Stn 41) 4. Pueblo, right emb. (Stn 46) 5. Belle Fourche (MP5)
6. San Luis (Stn 112) 7. San Luis (Stn 80) 8. Cairn Curran (R4) 9. Horsetooth (Stn 10) 10. Dixon Canyon (Stn 9)

Figure 7.87: Post construction lateral displacement of the upper upstream slope to
upstream crest region for selected case studies of embankment with earthfills and
gravels in the upstream shoulder.
Chapter 7: Deformation behaviour of embankment dams Page 7.158

7.7 GENERAL DEFORMATION B EHAVIOUR OF P UDDLE CORE


EARTHFILL EMBANKMENTS

7.7.1 DEFORMATION DURING CONSTRUCTION OF PUDDLE CORE EARTHFILL


EMBANKMENTS

Selset Dam (Section 5.6 of Appendix G) is the only puddle dam for which deformation
behaviour during construction has been found in the published literature. It was
constructed in the late 1950’s using relatively modern compaction techniques compared
to the remainder of the data set. Therefore, the deformation behaviour of Selset Dam is
not likely to be typical of most puddle dams. However, the deformation behaviour and
pore water pressure response during construction at Selset dam does highlight several
important aspects raised by Bishop and Vaughan (1962) fundamental to puddle cores:
• Arching develops across the core resulting in significant reductions in the total
vertical stresses within the core.
• The deformation of the core occurs predominantly as yielding in undrained loading
conditions, and the large observed settlements are primarily due to lateral
deformation of the narrow puddle core.
• Consolidation type settlements during construction are not significant.
• Relatively high pore water pressures are developed in the puddle core during
construction and are still present at the end of construction.
• The internal settlement of the puddle core during construction at Selset Dam (Figure
7.88) is well in excess of that observed in wet placed, rolled clayey earthfills in
zoned embankments. The vertical strains in Selset dam reached as high as 5 to 7%
at 10 m depth below crest level. In comparison, vertical strains at a similar depth
below crest were less than about 2% in zoned embankments with narrow rolled
earthfill cores (Figure 7.22).

The issue of arching in embankments with narrow cores is discussed in Section


7.4.1.3 with reference to the literature on puddle core earthfill embankments.
The rate of dissipation of pore water pressures in the puddle core is likely to be
affected by the permeability and width of the core, as well as the permeability of the
adjacent shoulder earthfill if it is similar to that of the puddle core.
Chapter 7: Deformation behaviour of embankment dams Page 7.159

The amount of lateral deformation of the puddle core will primarily be influenced
by the differential lateral stress conditions between the puddle core and supporting
shoulders and the compressibility of the supporting shoulder earthfill. The lateral
stresses that develop within the puddle core will depend on its shear strength and
compressibility properties, as well as the stress reducing influence of arching of the
core.

7%
Select fill (Gauge S1 to S2)
6% Select fill (Gauge S2 to S3)
Select fill (Gauge S3 to S4)
5%
Core settlement at end of construction
Vertical strain (%)

4%

3%

2%

1%

0%

-1%
0 5 10 15 20 25 30 35 40
Fill height above gauge (m)
Figure 7.88: Selset Dam internal settlements during construction (data from Bishop and
Vaughan 1962)

7.7.2 DEFORMATION DURING FIRST FILLING OF PUDDLE CORE EARTHFILL


EMBANKMENTS

Charles (1998) identifies the poorly compacted upstream shoulder of the older puddle
dams (presumably those embankments constructed prior to the 1900’s without formal
compaction) as prone to collapse compression on impounding.
Information on the susceptibility of earthfills to collapse compression from the case
study data on rolled earthfills is discussed in Section 7.5.2. In the context of puddle
dams, the susceptibility of materials to collapse compression and its effect on the
deformation behaviour of the embankment is summarised as follows:
• For puddle dams constructed prior to 1900, typically without any formal compaction
of the shoulder earthfill, collapse compression on inundation of the shoulders is
likely to be significant.
Chapter 7: Deformation behaviour of embankment dams Page 7.160

• For dams constructed without a select zone adjacent to the core, which was typical
before 1860, collapse compression is likely to be significant. From the available
literature it is apparent that the select zone adjacent to the core was placed in thinner
layers, thereby possibly reducing the amount of collapse compression adjacent to the
puddle core.
• For puddle core earthfill dams constructed after the late 1930s, when soil mechanics
principles were applied to embankment construction, moisture conditioning and
good compaction of the shoulder fill could be expected (e.g. Selset Dam). As a
result, collapse compression of the upstream shoulder earthfill on inundation is not
likely to have occurred.

The puddle core itself is not susceptible to collapse compression. However, because
of its very low undrained strength its post construction deformation behaviour is highly
dependent (or largely controlled by) the deformation behaviour of the supporting
shoulders. If the upstream (or downstream) shoulder region adjacent to the core settles
due to collapse compression on wetting, then the puddle core will settle with the
upstream shoulder. Therefore, the magnitude of crest settlement of the puddle core
embankment on first filling will depend on the embankment zoning geometry and the
susceptibility of the upstream shoulder earthfill zones to collapse compression on
wetting. The mechanism of deformation of the puddle core on first filling is similar to
that of zoned embankments with wet placed cores of low undrained strength and thin to
medium width (Section 7.5.2, Figure 7.35).
For puddle dams with well-compacted earthfills either side of the puddle core that
are not susceptible to collapse compression, crest settlements will be limited during first
filling as observed at Selset and Burnhope dams (Figure 7.92). In the case of Burnhope
dam (constructed in 1935) the upstream filling was placed in 450 mm layers and
compacted by 10 to 17 tonne steamrollers. In the first four years after construction the
settlement of the upstream slope was 104 mm (or 0.4% of the dam height) and 468 mm
(or 1.2%) at the crest. For Selset dam, where the shoulder filling was well compacted
with moisture content control, the crest settled 177 mm (0.45% of the dam height) on
first filling. As shown in Figure 7.92 the magnitude of crest settlement for these dams
is relatively small compared to other puddle core dams and the settlement versus time
curve shows virtually no influence due to first filling.
Chapter 7: Deformation behaviour of embankment dams Page 7.161

Where the shoulder fill adjacent to the puddle core is dry placed and poorly
compacted, and therefore susceptible to large magnitude collapse settlements on
wetting, crest settlements will be large on first filling as observed at Hope Valley dam
(Figure 7.92). The mechanics of the deformation behaviour for this condition are
discussed in more detail in Section 7.7.2.1.

7.7.2.1 Collapse Compression of Poorly Compacted Filling


The mechanism of deformation considered to be involved during first filling and first
drawdown of an old puddle dam with no formal compaction of the shoulder filling is
summarised as follows:
• On first filling:
− Wetting of the poorly compacted upstream filling results in collapse of the soil
structure and therefore large settlement of the upstream fill (Figure 7.89a to c).
The strength and compressibility properties of the wetted earthfill will be
significantly different to those of the “as placed” earthfill; its compressibility and
undrained strength will be significantly reduced.
− The effect of first filling on the deformation of the upstream shoulder is for large
settlement as a result of the collapse of the soil structure on wetting.
− At the upstream (and downstream) interface between the puddle core and
shoulder, the core must settle with the shoulder. This is because the puddle core
does not have the shear strength to support any significant levels of stress that
would be transferred to it by differential settlement.
− The total horizontal stress acting on the upstream face of the puddle core must at
least equal the lateral stresses within the puddle core for equilibrium.
Hydrostatic pressures acting on the upstream face of the puddle core increase
with the rise in reservoir level, but the lateral stresses within the upstream
shoulder decrease due to the substantial loss in strength and compressibility on
collapse due to wetting (Figure 7.89d). The deformation of the upstream face of
the core / shoulder interface is dependent on the net lateral stress acting. If the
net lateral stress at any point increases then the interface at that point will
deform downstream and horizontal stresses will increase in the puddle core and
downstream shoulder. If the net lateral stress decreases then the interface will
deform upstream until equilibrium conditions are re-established as lateral
Chapter 7: Deformation behaviour of embankment dams Page 7.162

stresses increase in the upstream shoulder. It is possible that the interface can
initially deform upstream at low water levels and then deform downstream as the
reservoir approaches full supply level.
− If upstream displacement of the core / shoulder interface occurs then lateral
spreading of the core results. The deformation of the puddle core occurs as
undrained plastic yielding and is therefore associated with vertical strain to
maintain volumetric consistency. Further arching across the core is also likely
where lateral spreading occurs, thereby decreasing the lateral stress required for
equilibrium.
− Once full supply level is reached, whether or not equilibrium pore water
pressure conditions in the upstream shoulder have been reached is dependent on
its permeability and the time for first filling. Free draining earthfills will
effectively reach equilibrium conditions as the reservoir is filled. For the case
studies analysed, in most cases the earthfill comprises a mix of clay to gravel
(and larger) sized particles and its permeability is quite variable due to layering
as well as material and density variations within and between layers. Therefore,
it is likely that for most cases equilibrium pore water pressure conditions are not
reached in the upstream shoulder when full supply level is reached (Figure
7.89b). As a result, the upstream shoulder will continue to wet up after first
filling is completed and settlements can continue to occur post first filling within
the shoulder and possibly the puddle core.
• On first drawdown:
− Further compression of the upstream shoulder occurs as effective stresses
increase under the falling reservoir level. Because the upstream earthfill was not
fully saturated at end of first filling, its drained compressibility will have
continued to decrease as further softening occurred on wetting post first filling.
Therefore, poorly compacted earthfills susceptible to collapse compression and
softening on wetting are likely to be close to normally consolidated after the end
of first filling, and increases in effective stresses on drawdown will result in
large deformations due to yielding. The upstream shoulder can therefore
experience large settlements on drawdown (Figure 7.90).
− The puddle core will also undergo significant settlement and displacement on
initial drawdown (Figure 7.90). As the hydrostatic pressures acting on the
Chapter 7: Deformation behaviour of embankment dams Page 7.163

upstream face of the puddle core decrease, total stresses in the puddle core and
downstream shoulder initially decrease and the embankment deforms upstream.
At some point, the lateral stresses in the upstream shoulder must increase to
maintain equilibrium with those within the puddle core. Because of the large
reduction in compressibility of the poorly compacted upstream shoulder earthfill
on wetting, lateral spreading of the puddle core will occur as the upstream
shoulder deforms under the increased lateral stresses imposed on it as
hydrostatic pressures continue to decrease on drawdown. The deformation of
the puddle core occurs as undrained plastic deformation (i.e. yielding) and the
lateral spreading is associated with vertical strains to maintain volumetric
consistency, and therefore crest settlements are potentially large.

In summary, deformation of the upstream shoulder earthfill (where it is poorly


compacted) will be significant on initial filling due to collapse compression. The
settlement of the puddle core will also be large on first filling as it deforms with the
adjacent upstream zone of the shoulder. On first drawdown large deformations can
occur in the upstream shoulder due to drained yielding as effective stresses increase in
the near normally consolidated water softened earthfill. Deformations of the puddle
core will be large on first drawdown mainly due to lateral spreading as the now softened
adjacent shoulder deforms under the increased lateral stresses applied as hydrostatic
pressures decrease. Subsequent drawdowns that result in increases in effective stresses
in the upstream shoulder above those previously experienced post first filling will also
result in large deformations of the upstream shoulder and puddle core (this is discussed
in Section 7.7.3.5).
In the case of wetting of part of the downstream earthfill shoulder (Figure 7.91),
such as due to seepage through the foundation, or hydraulic fracture through the puddle
core due in part to arching of the narrow core, the wetted zone of the filling would
undergo collapse compression assuming it to be in a poorly compacted and dry
condition. Consequently, deformations of the downstream shoulder and crest region are
likely to be large.
Chapter 7: Deformation behaviour of embankment dams Page 7.164

Figure 7.89: “Idealised” model of collapse compression of poorly compacted shoulder


fill of puddle core earthfill dam on first filling.

Figure 7.90: “Idealised” model of yielding of poorly compacted shoulder fill and puddle
core on first drawdown of a puddle core earthfill dam.
Chapter 7: Deformation behaviour of embankment dams Page 7.165

Figure 7.91: “Idealised” model of collapse compression on wetting of poorly compacted


earthfill in the downstream shoulder of a puddle core earthfill dam.

In the case of Hope Valley dam (refer Section 5.3 of Appendix G), wetting of the
downstream fill was observed shortly after the start of initial filling. This was probably
mainly due to seepage in the Tertiary alluvial soils in the foundation. From Figure 7.92
it is evident that significant crest settlement (1250 mm or 5.5% of the embankment
height) occurred in the period up to 11 years after construction. Collapse compression
on wetting of both the upstream and downstream earthfill shoulders were likely to be
significant contributing factors to the large crest settlement, in addition to the
deformations due to yielding on drawdown. What influence wetting and collapse
compression of the downstream shoulder earthfill had on the magnitude of crest
settlement is not known, and the data in Figure 7.92 does not include many comparable
dams. But, in comparison to the anecdotal information of the crest settlement for
several old UK puddle dams the magnitude of crest settlement of Hope Valley dam is
relatively high, possibly suggesting its influence was significant.
In the case of the Dale Dyke dam failure in 1864, Charles (1998) suspected that
collapse compression on wetting of the earthfill in the downstream shoulder might have
contributed to the loss of freeboard. Although, Binnie (1978) surmises that the large
settlement was localised to a small portion of the crest and due to loss of material from
piping, a view supported by Foster (1999) and Foster et al (2000). Instability of the
downstream slope prior to overtopping as a result of strength loss within the
downstream earthfill shoulder could also have been a factor.
It is notable that longitudinal cracking between the core and upstream filling has not
been reported on any puddle dams as a result of collapse compression, as has been the
case for a number of central core earth and rockfill dams. It is considered that this is
due to the low undrained strength and plastic nature of the puddle clay core. Rather
Chapter 7: Deformation behaviour of embankment dams Page 7.166

than a crack appearing, the core deforms with the upstream shoulder and may result in
the ridge or bulge developed on the upstream face near to crest level that it evident on
several puddle dams, or a narrow wedge of upstream zone adjacent to the core dropping
to form a “reverse” scarp as is the case for Happy Valley dam.

7.7.3 DEFORMATION BEHAVIOUR POST FIRST FILLING OF PUDDLE CORE EARTHFILL


DAMS

Section 7.2.3.3 summarises the research by BRE and Imperial College in terms of the
factors they consider to affect the long-term deformation behaviour and the
mechanism/s affecting the deformation behaviour of puddle core earthfill embankments.
The following sub-sections present an analysis of the deformation behaviour of the
puddle core earthfill dam case studies in the context of defining “normal” deformation
behaviour, and the various mechanisms associated with the observed deformation
behaviour are discussed. Section 7.12 discusses the methods of identification of
“abnormal” deformation behaviour of puddle core earthfill embankments and
summarises several case studies considered to indicate “abnormal” deformation
behaviour.

7.7.3.1 Available Deformation Records


The database comprises seventeen case studies of puddle core earthfill dams. The
embankment details, reservoir operation, hydrogeology and monitoring for each case
study are summarised in Table F1.5 in Section 1 of Appendix F. Table 7.24 presents a
summary of the post construction surface deformation behaviour of the crest and slopes
for each case study. Figure 7.92 presents the post construction crest settlement versus
log time for those embankments with records available from end of construction (time is
in years and zero time is the end of construction).
Figure 7.92 shows that the post construction crest settlement of puddle core earthfill
embankments is significant, ranging from 1% up to 8 to 14% of the dam height after
more than 100 years. It is generally the older (pre 1900) embankments that show the
greater post construction deformations. The more recent puddle core earthfill
embankments constructed in the 1930’s to 1950’s generally show significantly lower
crest settlement (less than 1 to 4% at 10 to 50 years after construction) as expected
Chapter 7: Deformation behaviour of embankment dams Page 7.167

given the differences in placement methods of earthfill and compaction equipment


between the two periods.

0%
Crest settlement (% of dam height) .

Selset
2%

Happy Valley Hope Valley


4%

Yan Yean
6% Hope Valley
Happy Valley
Selset
8%
Burnhope
Challacombe
10% Hollowell
Langsett
Ramsden
12%
end of first filling

14%
0.1 1.0 10.0 100.0 1000.0
Time (years since end of construction)
Figure 7.92: Post construction crest settlement versus log time of puddle core earthfill
dams.

The understanding of “normal” long-term deformation behaviour is mainly aimed at


three situations:
• Deformations during “steady state” conditions. This is basically deformation
behaviour assuming the reservoir level is maintained at a constant level, and
therefore the components of long-term deformation are foundation settlement,
shrink-swell movements, and primary creep type movements under constant stress
conditions. Clearly, the data set would be fairly limited given fluctuations in
reservoir level occur in most dams, however, as will be discussed later, it is
appropriate not only to dams with little fluctuation in reservoir level, but also to
cases where the embankment design incorporates a relatively impermeable zone
upstream of the core that effectively acts as a water barrier thereby limiting the
changes in the phreatic surface in the central portion of the dam.
• Deformations during the normal cyclic operation of the reservoir for puddle core
earthfill dams with permeable upstream earthfill.
• Deformations during a historically large drawdown event.
Chapter 7: Deformation behaviour of embankment dams Page 7.168

Table 7.24: Summary of the post construction surface deformations of the puddle core earthfill dam case studies.
Deformation of Embankment Crest Upstream Slope Downstream Slope Downstream Toe
Response to
Settlement since
Year Height, Reservoir Drawdown Period of Long-term Deformation Settlement
Name EOC Displacement Long-term Long-term Long-term Long-term
Completed H (m) Operation Upstream of Monitoring
Displacement Rate Rate (mm/yr) Settlement Displacement Settlement Displacement
Puddle Core mm % of H SLT SLT Monitored
(mm/yr)
assume 702 130 mm in 22 yrs
Burnhope 1935 41 assume minor EOC to 1957 1.8 0.80 - 0.14 - - - - -
fluctuating (22 yrs) since EOC
520
Challocombe 1944 15 Steady assume minor EOC to 1981 3.45 0.80 - - - - - - - -
(45 yrs)
negligible when at FSL
Cwnwernderi 1901 22 Steady assume full 1977 to 1989 - - 0.90 - - - - - - -
(> 70 yrs after EOC)

assume
Dean Head 1840 19 assume full 1977 to 1984 - - 7.4 - - - - - - - -
fluctuating
up to 2.5 mm/yr (85
2.3 to 0.7 to 102 yrs after 11 mm in 10 yrs
1 mm/yr over 10 5 mm over 10 0.6 mm/yr (92 to
Fluctuating (15 EOC to 1973 770 (varies with 29 mm (85 to 102 EOC), rate is (92 to 102 yrs
Happy Valley 1896 25 minor 3.1 1.0 - yrs (92 to 102 yrs yrs (92 to 102 102 yrs after
to 30% of H) 1982 to 1998 (76 yrs) magnitude of yrs after EOC) dependent on after EOC), SLT =
after EOC) yrs after EOC) EOC)
drawdown) magnitude of 1.0%
d'down.
assume 817
Hollowell 1937 13.1 assume minor EOC to 1954 6.2 0.70 - - - - - - - -
fluctuating (17 yrs)
assume 2.3 (only 18 months negligible (150 yrs after
Holmestyes 1840 25 negligible 1990 to 1991 - - - - - - - - -
fluctuating data) EOC)
5 mm/yr at 118 to 128
negligible on mid 0.5 mm/yr (118 to
Fluctuating (15 variable (minor to EOC to 1983 1820 yrs after EOC
Hope Valley 1872 21 8.7 2.6 - - - slope (118 to 128 128 yrs after - -
to 25% of H) full) 1988 to 1999 (125 yrs) (downstream edge of
yrs after EOC) EOC)
crest)
assume
Ladybower 1945 43 assume minor 1950 to 1984 - - 2.0 - - - - - - - -
fluctuating
assume 124
Langsett 1904 33 nk EOC to 1907 3.8 - - - - - - - - -
fluctuating (3 yrs)
10 (1 normal 7 mm on 10 m 4 mm on 10 m
assume
Ogden 1858 25 assume full 1989 to 1993 - - drawdown cycle - - d'down cycle in - d'down cycle in - - -
fluctuating
only) 1992/93 1992/93

12 mm in 2 yrs
(450 mm 6.6 (normal 8 mm in 2 yrs -2 mm in 2 yrs
assume 1949 to 1985 (1988 to 1990), -8 mm in 2 yrs (1988 2 mm in 2 yrs negligible (1988
Ramsden 1883 25 assume full from 1949 to - operating 3 mm/yr (1977 to 1985) - (incl. Extreme (incl. Extreme
fluctuating 1988 to 1990 includes abnormally to 1990) (1988 to 1990) to 1990)
1985) conditions) d'down) d'down)
large d'down

assume 313
Selset 1959 39 minor EOC to 1970 0.80 0.4 - - - - - - - -
fluctuating (10 yrs)

some early 2 mm on 13 m 5 to 6 mm on 13
Fluctuating (50 4.5 (1915 to 1943) 3 mm/yr (85 years after < 1 mm (1988 to 1 to 2 mm (1988
Walshaw Dean 1907 22 assume full data, 1990 to - - 1.7 d'down in 1988 to m d'down cycle in - -
to 60% of H) to 6.7 (1988/90) EOC) 1990) to 1990)
1995 1990. 1988/90

assume
Widdop 1878 20 assume full 1988 to 1990 - - - - - - - - - - -
fluctuating

negligible rate at
15 to 20 mm in 13
0.5 (upstream edge upstream edge of crest. max 8 - 10 mm/yr,
2 to 3 mm/yr negligible yrs (more than
Fluctuating (15 1205 of the crest) Up to 4.5 to 5.5 mm/yr 4 mm in 6 yrs (> negligible (135 to increasing rate
Yan Yean 1857 9.6 minor 1986 to 1999 12.5 2.0 (more than 128 (more than 130 130 yrs after
to 25% of H) (128 yrs) 11 (downstream (increasing rate) on 135 yrs after EOC) 141 yrs after EOC) with time (> 128
years after EOC) yrs after EOC) EOC), stick - slip
edge of the crest) downstream crest and years after EOC)
type displacement
slope -

5.4 (120 yrs after


Yateholme 1872 17 Fluctuating assume full 1989 to 1992 - - - - - - - - - -
EOC)
Chapter 7: Deformation behaviour of embankment dams Page 7.169

Notes to Table 7.24:


SLT = long-term settlement index in unit of settlement (as a percentage of dam height) per log
cycle of time (refer Equation 7.6 in Section 7.7.3.2).
EOC = end of construction FSL = full supply level
Displacement is the deformation in the horizontal direction normal to the dam axis, positive is
downstream and negative is upstream.
d’down = drawdown

In Table 7.24 the reservoir operation for each case study is categorised as either
fluctuating or steady, and the level of fluctuation, where it can be estimated, is given as
a percentage of the embankment height. For those case studies where the reservoir
operation is not known, it has been assumed as fluctuating. Also presented in Table
7.24 is an estimate of the response of the phreatic surface in the zone upstream of the
puddle core to changes in reservoir level (more details are given in the Table F1.5 in
Section 1 of Appendix F). The assessment is based on piezometer records where
available, but in their absence has been assessed from the presence or otherwise of a
select earthfill zone upstream of the core. The purpose of categorising the response of
the phreatic surface in the zone upstream of the puddle core is because it has important
implications on the deformation behaviour (as discussed in Sections 7.7.3.2 to 7.7.3.6).
Three categories of response have been used:
• Full or large pore water pressure response to reservoir fluctuation, indicating the
earthfill in the upstream shoulder to be permeable.
• Minor response, indicating the earthfill zone upstream of the puddle core is of low
permeability such that the change in pore water pressure recorded is a small
percentage of the change in reservoir level.
• Negligible response, indicating virtually no change in pore water pressure in the
earthfill zone upstream of the puddle core due to the reservoir operation. Only
Holmestyes dam is in this category because it has an effective puddle clay blanket
on the upstream face of the dam.

It should be noted that the categorisation of the response of the phreatic surface
upstream of the puddle core to changes in reservoir level is based on a limited
knowledge of the reservoir operation and assumptions regarding the permeability of
earthfill zones in the upstream shoulder. For embankments constructed after 1930 it has
been assumed that the select filling is well compacted and of low permeability. For
Ramsden, Walshaw Dean (Lower), Ogden and Yateholme dams (all Yorkshire Water
Chapter 7: Deformation behaviour of embankment dams Page 7.170

dams) the shoulder earthfill comprised weathered Millstone Grit and has been assumed
as permeable based on the results of constant head tests at Ramsden dam (permeability
varying from 2 × 10-5 m/sec to 10 -7 m/sec). For the Australian dams Yan Yean, Hope
Valley and Happy Valley the classification has been based on the pore water response
from piezometers installed in the upstream shoulder.
It is recognised that the compressibility of the zones supporting the puddle core have
a significant influence on the deformation behaviour of the core (and settlement of the
crest) under cyclic reservoir conditions. The compressibility of the earthfill shoulders
will generally be lower and not as susceptible to softening on wetting where it has been
well compacted (i.e. post 1930) compared to poorly compacted earthfills. Material type
will also influence the material compressibility.

7.7.3.2 Long-term Settlement Rate Index


The long-term settlement rate index, SLT, is used as a guide for prediction of settlements
(mainly crest settlement) of puddle core earthfill dams post first filling and as a
comparative measure to assess the performance of a specific dam. The settlement rate
index, S LT, is calculated using Equation 7.6. It is the same equation as the settlement
index proposed by Charles (1986) except that the settlement rate is expressed in units of
settlement (as a percentage of dam height) per log cycle of time. The same form of
equation has been used to calculate the post first filling long-term settlement rate for the
crest and shoulders of rolled earth and earth-rockfill embankments.

S LT = s ∗ 100 /[1000 ∗ H ∗ log(t2 t 1 )] (7.6)

where s is the crest settlement in millimetres measured between times t2 and t1 (in years)
after the end of embankme nt construction, and H is the height of the dam in metres.
SLT values for the puddle core earthfill dam case studies are given in Table 7.24 and
have been calculated mainly for the crest of the embankments. Some values are given
for the upper regions of the upstream and downstream shoulders. The quoted S LT values
are for periods of measurement during normal operation conditions (i.e. excluding
abnormally large drawdown events) and are generally from deformations observed tens
of years after construction.
The long-term settlement rate, SLT, basically encompasses all of the mechanisms
involved in the long-term deformation behaviour (summarised in Section 7.2.3.3) into
Chapter 7: Deformation behaviour of embankment dams Page 7.171

one factor appropriate to analysis of crest settlement for periods of normal reservoir
operating conditions. Analysis of the data indicates that a correlation exists between SLT
for crest settlement, the reservoir operation (under normal conditions) and the pore
water pressure response of the earthfill zone upstream of the puddle core to changes in
reservoir level (i.e. full, minor or negligible as discussed in Section 7.7.3.1). The
correlation is summarised as follows:
• For conditions where the change in pore water pressure in the earthfill zone
upstream of the puddle core is a small fraction of the change in reservoir level on
drawdown or where the reservoir remains steady, SLT generally ranges from 0.4% to
1% settlement per log cycle of time (settlement as a percentage of the embankment
height).
• For fluctuating reservoir conditions and where the change in pore water pressure in
the earthfill zone upstream of the puddle core closely follows the reservoir level, SLT
ranges from 2.6% to 7.4% settlement per log cycle of time.

7.7.3.3 “Normal” Deformation Under “Steady State” Conditions

The results of long-term post construction monitoring indicate the reservoir operation
and the pore water pressure response of the earthfill zone upstream of the puddle core to
changes in reservoir level have a significant influence on the long-term deformation
behaviour. Case studies of “normal” deformation behaviour under “steady state”
conditions (Figure 7.93) are consider to comprise those cases where the reservoir level
is essentially maintained at a steady level (Cwnwerderi and Challocombe dams), or the
pore water pressure response of the earthfill zone upstream of the puddle core is either
negligible (Holmestyes) or minor (Burnhope, Hollowell, Ladybower, Selset, Yan Yean
and Happy Valley dams) to changes in the reservoir level.
For these nine case studies the long-term crest settlement rate, S LT, typically ranges
from 0.4% to 1% settlement per log cycle of time.
Of these embankments both Ladybower and Holmestyes dams exceed an SLT of 1%.
For Holmestyes (SLT of 2.3%) the period of monitoring of 1.5 years at 150 years after
construction is considered too small to obtain a reasonable estimate, as only 3 mm of
crest settlement occurred during this period. Ladybower dam (SLT of 2.0%) was
classified as “minor” response to reservoir level fluctuation based on the year of
completion of construction (1945) and assumption that a zone of select earthfill of low
Chapter 7: Deformation behaviour of embankment dams Page 7.172

permeability was used up and downstream of the puddle core, as was the general design
practice used at the time. However, the embankment design details, the construction
materials used and their method of placement were not known, and it is therefore
possible that these assumptions are not correct for Ladybower dam.

Figure 7.93: Definitions of “steady state” conditions during normal reservoir operating
conditions.

Of the remaining seven cases considered as “steady state” conditions there is no clear
correlation of the long-term crest settlement rate, SLT, with embankment height or age.
However, SLT is be expected to be affected by:
• The age of the embankment, mainly due the difference in design and compaction
techniques adopted at the time of construction. The S LT of the older puddle dams
would be expected to be greater than for newer dams constructed after the 1930’s.
• The height of the embankment. It is not clear what effect the embankment height
has because its influence appears to be relative minor and over-shadowed by other
factors, as is the case with the earth and earth-rockfill embankments (Figure 7.62
Chapter 7: Deformation behaviour of embankment dams Page 7.173

and Figure 7.63). It would be expected that SLT would be greater for dams of greater
height.
• Pore water pressure changes in the embankment. Even though steady state
conditions have been assumed, the pore water pressures in the puddle core and the
zone immediately upstream would be expected to change with time. The available
data does indicate some change, albeit minor, with reservoir level.
• The material type and placement method of the various zones within the
embankment, and their effect on the time dependent creep coefficients. The data for
rolled earthfill and earth-rockfill embankments (Section 7.6.3) indicates clayey
earthfills generally have higher long-term settlement rates than dominantly sandy
earthfills, and dams with poorly compacted rockfill shoulders higher long-term
settlement rates than well-compacted rockfills.

In summary, it is considered based on the available records of the case studies


analysed that under “steady state” conditions (or near steady state conditions) the
expected long-term crest settlement rate, SLT, for “normal” deformation of puddle core
earthfill dams is less than about 1% settlement per log cycle of time. The mechanisms
involved are considered to be time dependent primary creep and deformations
associated with the small changes in effective stress following the small changes in pore
water pressure in the puddle core and upstream shoulder.
Holmestyes Dam (constructed 1840) presents a slightly different case to most
puddle core earthfill embankments classified as “steady state” conditions. A layer of
puddle clay was placed on the upstream face of the dam in 1857 (some 17 years after
construction) due to the observations of leakage (Tedd et al 1993). Installation of
piezometers downstream of the upstream blanket indicated the phreatic surface in the
upstream shoulder and core to be well below the reservoir level, and on drawdown only
minimal responses in piezometric level were recorded. On a large drawdown of
approximately 10 m, minor heave movements of the crest and upstream shoulder were
observed (Tedd et al 1993).
Chapter 7: Deformation behaviour of embankment dams Page 7.174

7.7.3.4 “Normal” Deformation Under Normal Reservoir Operation and Fluctuating


Phreatic Surface in the Upstream Shoulder

This section looks at case studies where large changes in pore water pressure occur in
the upstream shoulder (adjacent to the core) with fluctuations in reservoir level (Figure
7.94). In these cases cyclic changes in effective stress occur in the upstream shoulder.
Cyclic changes in total stress acting on the upstream face of the puddle core also occur
as hydrostatic pressure change with fluctuations in reservoir level. Long-term
settlement rates of the crest and upstream slope for these cases are generally
significantly greater than for “steady state” conditions as described in Section 7.7.3.3.
The case studies include Dean Head, Ramsden, Yateholme and Walshaw Dean
dams where the long-term crest settlement rates, SLT, under normal operating conditions
range from 4.5 to 7.4 % per log cycle of time (Table 7.24). The rate of crest settlement
of these dams is typically 5 to 10 times that under “steady state” conditions and
confirms the significant influence of the permeability of the earthfill zone upstream of
the puddle core under fluctuating reservoir conditions on the long-term crest settlement
rate. Figure 7.92 graphically presents the difference in the long-term crest settlement
rate between Ramsden dam and several “steady state” dams.

Figure 7.94: Fluctuation of pore water pressures adjacent to and upstream of the puddle
core with fluctuations in reservoir level.

For Hope Valley dam (SLT of 2.6%) the pore water pressure response in the earthfill
zone upstream of the puddle core is variable, ranging from 25% to 100% of the change
in reservoir level. It has been included with this group of case studies even though it
possibly lies somewhere between these case studies where large changes in pore water
pressure upstream of the puddle core are observed and the “steady state” case studies.
The reported long-term crest settlement rate for Hope Valley dam is from SMPs on the
Chapter 7: Deformation behaviour of embankment dams Page 7.175

downstream edge of the crest, and the crest settlement rate above the puddle core is
therefore possibly higher than 2.6% per log cycle time.
Ladybower dam possibly also falls into this category, however, no information was
available on the pore water pressure response under drawdown or the reservoir
operation to confirm or deny this.
The mechanism of deformation during the normal cyclic operation of the reservoir
for embankments with permeable upstream earthfill is:
• On drawdown:
− Increasing effective vertical stresses in the upstream shoulder occur as pore
water pressures dissipate relatively quickly in the permeable earthfill, resulting
in settlement. Settlements are not expected to be overly high because the stress
path is in the unloading / reloading range and the earthfill’s compressibility will
therefore be relatively low.
− The reduction in hydrostatic pressure acting on the upstream face of the puddle
core results in a reduction in the total horizontal stresses within the puddle core
and downstream shoulder, and upstream displacement occurs. This upstream
displacement was observed for SMPs on the crest, upstream slope and upper
downstream slope of Walshaw Dean and Ramsden dams (refer Section 5 of
Appendix G) and the internal displacement of the centre of the puddle core at
Ramsden dam (Figure 7.95) during drawdowns within the normal operating
range for the reservoir.
− Some of the reduction in hydrostatic pressure acting on the upstream face of the
puddle core is transferred onto the upstream shoulder in order to maintain
equilibrium with the lateral stresses of the puddle core. This increase in
horizontal stress in the upstream shoulder will result in a net upstream
displacement of the upstream interface of puddle core and lateral spreading of
the puddle core. Deformations within the puddle core occur primarily as
undrained plastic yielding and therefore lateral spreading is accompanied by
vertical strain to maintain volumetric consistency, and therefore settlement of the
crest. Internal vertical strains measured in the puddle core during normal
operating drawdown at Ramsden (Figure 7.96) and Walshaw Dean dams
confirm this pattern of behaviour (refer Sections 5.4 and 5.5 of Appendix G).
Chapter 7: Deformation behaviour of embankment dams Page 7.176

− Vertical strains in the puddle core can also occur due to increases in effective
vertical stress as a result of pore water pressure dissipation. The amount of pore
water pressure dissipation will depend on the permeability of the puddle core
and the time the reservoir is in a drawn down condition.

Figure 7.95: Ramsden Dam, internal horizontal deformation of the puddle core, 1988 to
1990 (Holton 1992).

• On re-filling the stress paths are reversed and:


− Decreases in effective vertical stresses in the upstream shoulder result in heave
type deformations.
− Increase in the horizontal stresses acting on the upstream face of the puddle core
results in downstream displacement of the core, upstream slope and downstream
slope.
Chapter 7: Deformation behaviour of embankment dams Page 7.177

− Within the puddle core heave type internal strains and heave of the crest are
observed (Figure 7.96). The mechanism causing this will be a combination of
reduction in lateral width of the core under the increased total stress acting (i.e. a
plastic type response on re-filling) and heave due reduction in effective vertical
stresses as pore water pressures increase. Deformations due to pore water
pressure increase will show a time delay due to the low permeability of the core.

0.7% 0
S3 S6
S4
0.6% 5

Reservoir Level (m below crest) .


S5
0.5% 10
S2
Vertical strain (%) .

0.4% 15

S1
0.3% 20

0.2% 25

0.1% 30

Crest to 6.5 m depth 6.5 to 9.1 m below crest


0.0% 9.1 to 12.6 m below crest 12.6 to 18.5 m below crest 35
from 18.5 m below crest reservoir
-0.1% 40
105.0 105.5 106.0 106.5 107.0 107.5
Time (years since end of construction)
Note: S1 to S6 refer to the stages of reservoir level
Figure 7.96: Ramsden Dam, internal vertical strains measured in the puddle core from
1988 to 1990 (data from Tedd et al 1997b and Kovacevic et al 1997).

The net permanent vertical deformations following a reservoir cycle of normal


magnitude are typically settlement of the crest, upstream slope and upper portion of the
downstream slope. The magnitude of settlement is generally greatest for the crest. Net
permanent crest displacements are generally small and in a downstream direction, but
data records are limited. Tedd et al (1997a) attribute the net permanent settlement of the
upstream shoulder to the difference in compressibility properties between unloading
(lower compressibility on re-filling) and re-loading (higher compressibility on
drawdown) as observed in cyclic laboratory oedometer testing (Holton 1992; Tedd et al
1997a). Kovacevic et al (1997) obtained good agreement between measured and
predicted settlements (from finite element analysis) under normal reservoir drawdown
cycle for several puddle core earthfill embankments when incorporating the differential
compressibility between unloading and re-loading in their analysis.
Chapter 7: Deformation behaviour of embankment dams Page 7.178

The net internal deformation of the puddle core at Ramsden dam shows upstream
displacement (Figure 7.95) and permanent vertical strains (Figure 7.96) over the full
height of the core. At Walshaw Dean dam the internal vertical strain shows a
permanent increase in strain over the full height of the core (refer Section 5.5 in
Appendix G). These permanent core deformations are most likely primarily due to
undrained plastic deformations on drawdown.
The net permanent crest settlements (Figure 7.97) are typically 50 to 70% of the
maximum settlement at drawdown, for drawdowns within the range of normal reservoir
operation. Permanent settlements of greater magnitude are observed for abnormally
large drawdowns, and are discussed in Section 7.7.3.5. The data in Figure 7.97 is
mainly from puddle dams in Yorkshire, England where the shoulder earthfill is
relatively permeable.
The permanent crest settlement on drawdown for puddle core earthfill dams with
permeable earthfill in the upstream shoulder is likely to be affected by the magnitude of
the drawdown, the length of time of the drawdown and the compressibility of the
upstream filling. Ignoring the events associated with abnormally large drawdown
events, Figure 7.97 shows some correlation of permanent crest settlement with
drawdown height. It should be noted that the data set is relatively small and comprises
UK puddle dams of similar height and constructed using similar materials for the
shoulders. Therefore, the following preliminary predictive methods, which are based on
the data presented by Tedd et al (1997b), should be used with caution:
• Estimation based on the SLT rates of the reported cases within the data set, ranging
from 4.5 to 7.4% vertical strain per log cycle of time.
• Estimation based on the plot of permanent crest settlement versus depth of
drawdown (as a percentage of the embankment height) as shown in Figure 7.98.
This plot is based on the data from Tedd et al (1997b) omitting those points
associated with abnormally large drawdown events and normalising the drawdown
height with respect to embankment height.

Where the pore water pressure response of the earthfill zone upstream of the core is
in-between that of “permeable” and low permeability, such as at Hope Valley dam (and
possibly Ladybower dam) the long-term settlement rate is also between that of the two
categories. However, there is not sufficient data with which to provide a possible range
of long-term crest settlement rate.
Chapter 7: Deformation behaviour of embankment dams Page 7.179

Figure 7.97: Vertical strain (or settlement) at the crest versus drawdown depth for
specific drawdown and refilling cycles (Tedd et al 1997b).

0.05%
Permanent crest settlement, .
Sp (% of dam height) .

0.04%

0.03%

0.02%

0.01%

0.00%
0 10 20 30 40 50 60 70 80
Drawdown (% of embankment height)

Figure 7.98: Permanent vertical crest strains during cyclic drawdown (in normal
operating range) for puddle core dams with permeable upstream filling (after Tedd et al
1997b).

7.7.3.5 Deformation During Abnormally Large Drawdown Events

Tedd et al (1997b) report the deformation behaviour during historically large drawdown
events for several UK puddle core dams, including Ramsden, Ogden and Widdop dams.
Permanent crest settlements under abnormally large drawdown for these three dams
were:
Chapter 7: Deformation behaviour of embankment dams Page 7.180

• 51 mm at Ramsden dam (Figure 7.99), or 0.20% of the embankment height, as a


result of the abnormally large drawdown cycle of 17 m at 106 years after
construction. Further details are given in Section 5.4 of Appendix G.
• 130 mm at Ogden dam (Figure 7.100), or 0.52% of the embankment height, during
the abnormally large 20 m drawdown event in 1990 to 1992.
• 52 mm at Widdop dam, or 0.26% of the embankment height, during the abnormally
large drawdown cycle of 17 m at 110 years after construction.

The data records indicate large permanent crest settlements occur under abnormally
or historically large drawdown events. At Ramsden dam a smaller magnitude, but still
relatively large, permanent settlement was measured for the SMP on the upstream slope.
The initial deformation pattern on drawdown to a level commensurate with the
normal operating conditions would result in the deformations as described in Section
7.7.3.4. On drawdown to abnormally or historically low levels, drained yielding in the
upstream shoulder and undrained yielding in the puddle core are considered to be the
dominant mechanisms resulting in the large magnitudes of deformation observed.

80 0

5
70

Reservoir Level (m below crest) .


10
60
15
Settlement (mm)

50
20
SMP U2 - upstream slope
40 SMP M - crest 25
SMP A - downstream slope
SMP G - downstream toe 30
30 reservoir
35
20
40

10
45

0 50
105.0 105.5 106.0 106.5 107.0 107.5
Time (years since end of construction)
Figure 7.99: Ramsden Dam, settlement of SMPs versus time from 1988 to 1990 (data
from Tedd et al 1990).
Chapter 7: Deformation behaviour of embankment dams Page 7.181

200 0

160 10
Settlement (mm)

Reservoir Level (m)


15

120 20

25

80 30

Ogden - crest 35

reservoir
40 40

45

0 50
131 132 133 134 135 136
Time (years since end of construction)

Figure 7.100: Ogden Dam, crest settlement versus time (Tedd et al 1997a).

In Section 7.7.2.1 it was considered that poorly compacted earthfills in the upstream
shoulder, susceptible to collapse compression on wetting, were close to normally
consolidated after the end of first filling due to the softening effects of collapse and
wetting. Increases in effective stresses on drawdown post first filling were considered
to result in large deformations due to yielding as vertical stress levels exceeded the pre-
consolidation pressure in the near normally consolidated upstream earthfill. On
subsequently larger drawdowns where the effective stresses in the upstream shoulder
exceeded those previously experienced post first filling, further large deformation of the
upstream shoulder due to yielding in drained conditions was also likely. It is on these
abnormally or historically large drawdown many years after the end of construction that
effective vertical stresses in the permeable upstream shoulder are considered to increase
to levels greater than previously experienced. The new high effective vertical stress
levels will be experienced in the lower elevations of the upstream shoulder (Figure
7.101) and drained yielding will be confined to this region as stress levels exceed the
pre-consolidation pressure and the earthfill’s compressibility increases significantly
above that for the re-compression range of loading. In the mid to upper regions of the
upstream shoulder it is likely that effective stress levels will equal those previously
experienced and therefore yielding will not occur; smaller settlements are likely under
the re-compression compressibility properties of the earthfill.
Chapter 7: Deformation behaviour of embankment dams Page 7.182

Figure 7.101: Idealised model of deformation during historically large drawdown.

For the puddle core, the hydrostatic pressures acting on the upstream face of the puddle
core will be reduced to levels not previously experienced. Therefore, the horizontal
stress levels in the upstream shoulder required to equilibrate the lateral stress from the
puddle core will exceed those levels previously experienced since first filling. The new
high lateral stress levels will be in the lower region of the embankment where a net
upstream displacement of the core / upstream shoulder will occur. Deformations in the
puddle core will largely occur as undrained plastic lateral spreading and vertical
compression (Figure 7.101). This pattern of deformation behaviour of the puddle core
was observed at Ramsden dam where vertical strains in the mid to lower region were
much greater than in the upper region (Figure 7.96) and the bulk of the lateral
deformation was below about 15 m (Figure 7.95).
After the initial plastic deformation, consolidation type settlements within the
puddle core may occur as pore water pressures dissipate. For Ramsden dam, Figure
7.95 and Figure 7.96 show significant lag in the deformation of the lower portion of the
core. It is suspected that this may not only due to pore water pressure dissipation in the
puddle core, but may also due to a lag in pore water pressure dissipation in the upstream
shoulder.
On-refilling after the historically large drawdown at Ramsden dam, a large amount
of the lateral displacement of the core was recovered (Figure 7.95) as the hydrostatic
pressures and horizontal stresses acting on the upstream face of the puddle core
increased. However, only a limited amount of the vertical strain was recovered on re-
filling (Figure 7.96). At Ogden dam (Figure 7.100) only a small proportion of the
maximum crest settlement (approximately 6%) was recovered on re-filling. The reasons
for this are considered to be due to the large undrained plastic type deformation of the
core and drained yielding in the lower upstream shoulder that is not recovered on re-
filling.
Chapter 7: Deformation behaviour of embankment dams Page 7.183

Comparison of the crest settlement of Ramsden (Figure 7.99) and Ogden (Figure
7.100) dams with Walshaw Dean Dam (Figure 7.102) during a full drawdown provides
some degree of verification of the mechanism discussed. As indicated, for both
Ramsden and Ogden dams the magnitude of crest settlement was large on full
drawdown, 51 mm and 130 mm respectively. In contrast, the full drawdown of
Walshaw Dean dam over the period 86 to 88 years after construction resulted in a
permanent crest settlement of only 9 mm, which is in line with the general settlement
trend from the earlier drawdowns. The reason for the deformation behaviour at
Walshaw Dean is considered to be because the normal reservoir operation is for very
large seasonal drawdown, as shown in Figure 7.102, and the maximum effective
stresses on full drawdown within the upstream earthfill have previously been
experienced. Therefore, the changes in stress are within the re-compression zone for
which the compressibility of the earthfill is significantly lower.

70 0

5
60
10

Reservoir Level (m below crest) .


50 15
Settlement (mm) .

20
40
25
30
30

20 35

SMP M1 - dam crest 40


10 reservoir
45

0 50
82 83 84 85 86 87 88 89
Time (years since end of construction)
Figure 7.102: Walshaw Dean (Lower) dam, crest settlement versus time (Tedd et al
1997a).

In terms of prediction of deformation during abnormal drawdown events, the only


available method would seem to be fully coupled finite element modelling (FEM) due
to the complexity of the processes and interaction between the different elements within
the embankment. Kovacevic et al (1997) attempted to model the deformation of
Ramsden dam during the large drawdown event using finite element methods, but were
unable to model the deformation with any accuracy. It is considered that they may not
Chapter 7: Deformation behaviour of embankment dams Page 7.184

have been able to model the yielding that occurred in the core and upstream zone when
the reservoir was drawn down to an all time low level. Therefore, undertaking FEM is
not a simple procedure and to add to the complexity is the lack of known information on
material properties and required assumptions necessary in the modelling.

7.7.3.6 Factors Affecting the Long-term Deformation Behaviour of Puddle Core


Dams
In summary, from the above analysis and discussion on the “normal” deformation
behaviour of puddle dams, the factors affecting the long-term deformation behaviour are
considered to be:
• Under normal reservoir operating conditions the permeability of the earthfill zone
upstream of the puddle core has a significant effect on the long-term deformation
behaviour of puddle core earthfill embankments. Long-term crest settlement rates,
SLT, under “steady state” conditions (or near steady state conditions) are typically
less than about 1 % per log cycle of time (settlement as a percentage of the
embankment height). For embankments where the earthfill upstream of the puddle
core is permeable and the seasonal reservoir fluctuation is large, the SLT rates are
significantly higher (2.6 to 7.5 % per log cycle of time).
• The historical operation of the reservoir also has a significant effect on the
deformation behaviour. Historically large drawdown events result in large,
permanent undrained plastic deformations of the puddle core, and yielding (and
possibly undrained plastic deformation) of the lower portion of the upstream filling.
This type of deformation behaviour could be classified as “abnormal”.
• The method of placement and moisture content control of the shoulder filling,
particularly, the upstream shoulder. The poorer the quality of compaction and drier
the moisture content at placement, the greater the potential for collapse compression
on saturation. The shoulder earthfill in the older (pre 1900) puddle core earthfill
embankments were likely to be particularly susceptible to collapse compression
given their generally poor compaction.
• Primary creep (or creep under constant stress), secondary consolidation and
foundation compression are encapsulated in the long-term settlement rate. Given
the age of most puddle core earthfill embankments these effects are considered to be
relatively minor compared with the effect of cyclic changes in stress in the upstream
Chapter 7: Deformation behaviour of embankment dams Page 7.185

earthfill zone and on the upstream face of the puddle core in embankments with
permeable upstream earthfill shoulders.
• Shrink-swell movements are likely to influence those sites for which plastic clays
are present and seasonal moisture content changes occur.
• Seasonal creep movements of the outer downstream slope. This is creep of the outer
surface of the downstream slope due to seasonal factors such as moisture content
profile changes. It is dependent on the plasticity of the soil and potential for shrink-
swell related movements, the steepness of the slope and the soil strength. The effect
of seasonal creep movements is usually offset in the method of installation of
surface measurement points on the slopes of the embankment.

7.8 “ABNORMAL” EMBANKMENT DEFORMATION BEHAVIOUR –


M ETHODS OF IDENTIFICATION
“Abnormal” deformation behaviour of embankment dams is deformation that cannot
readily be explained as being due to “normal” mechanisms. The purpose of attempting
to identify abnormalities in the deformation behaviour is for the early detection of
potential problems such as instability or internal erosion. Identification of “abnormal”
deformation behaviour is basically from outliers to the “normal” deformations trends
from similar embankment types taking into consideration the embankment zoning
geometry, material types, placement methods and foundation conditions. Outliers are
identified from magnitude of deformations as well as trends in the direction and rate of
deformation. Often what may initially be termed “abnormal” may later be proven by
investigation, analysis and additional monitoring to in-fact be “normal”.
During construction, identification of “abnormal” deformation is from outliers to the
total settlement and vertical strain profiles in the core of earthfill and zoned earth and
earth-rockfill embankments (Section 7.4.3). The data on lateral core deformations in
zoned earthfill embankme nts with thin to medium core widths (Section 7.4.2) from the
case studies is limited and therefore not sufficient for assessment of “abnormal”
deformation unless measured core deformations are larger than or the trend different to
the case study data presented.
Deformation of SMPs on the shoulders of embankments during construction is
available for some embankments in the literature, but it has not been considered here
due to inconsistencies in timing and variation in embankment type for which the data is
Chapter 7: Deformation behaviour of embankment dams Page 7.186

available. Penman (1986) provides some guidelines for the magnitude of lateral
displacement, but the data is from a variety of embankment types and should be used
with caution.
Identification of “abnormal” deformation post construction for earthfill and zoned
earth and earth-rockfill embankments is from outliers in terms of magnitude and rate of
deformation of surface monitoring points on the crest and slopes of the embankment
taking into consideration the zoning geometry and material properties of the core and
shoulders. “Abnormal” trends in the deformation behaviour, such as increased rates of
deformation on first filling, and internal localised zones of high strain in the core are
other means of identification of potentially “abnormal” deformation behaviour. For
puddle core earthfill dams outliers are more readily evaluated from the rate and trend of
the long-term deformation behaviour given the age and period of monitoring of these
structures.
It is important to note that for embankments where the deformation is considered to
be “abnormal” may be, and in most instances is, explainable on consideration of the
type and placement method of the earth or rockfill, or the reservoir operation. For
example, embankments with high vertical core strains during construction may be
related to plastic core deformations due to low undrained strength of the core and
corresponding large lateral strains of the supporting poorly compacted rockfill
shoulders. Whilst the core strain may be large in comparison to similar embankments
and therefore identified as “abnormal”, it may in-fact be “normal” given the reasons for
the behaviour.
It is emphasised that the identification of a dam as exhibiting “abnormal”
deformation behaviour in no way indicates there is any issue of dam safety with that
dam.
In the following sections (Sections 7.9 to 7.12) various methods of identification of
“abnormal” deformation behaviour are discussed with reference to and specific
discussion on those case studies within which the “abnormal” deformation was
observed.
The “abnormal” deformation behaviour may be specific to one part of the
embankment whilst the overall stability of the embankment is not in question. For
example, movement along a shear plane in the core on a large drawdown event for a
zoned earth and rockfill dam would be considered as “abnormal” behaviour of the core,
but the upstream rockfill zone provides adequate overall stability of the embankment.
Chapter 7: Deformation behaviour of embankment dams Page 7.187

Where failures due to slope instability have occurred in embankment dams there is
often limited, if any, information of the deformation behaviour leading up to failure. Of
the embankments considered in this chapter only at Carsington, Belle Fourche, San Luis
and Steinaker dams did a slope failure occur, and only at Carsington was there any
formal monitoring from within the failed region prior to the failure.
Chapter 5 presents information on the characteristics of pre and post failure
deformation behaviour from analysis of some 54 case studies of slope failure in
embankment dams with discussion on the mechanism causing the deformation
behaviour. Most of the data presented relates to the post failure deformation behaviour,
but there is some useful information on deformation behaviour leading up to failure that
is supplementary to the several case studies discussed within this chapter.
More generally, the basic concept of the creep model under constant deviatoric
stress conditions (Singh and Mitchell 1968; Mitchell 1993) is useful for assessment of
the deformation behaviour leading up to a failure condition. Primary creep, or
deformation at a decreasing rate with time under constant stress conditions, is applicable
to “normal” type deformation behaviour. Tertiary creep, or deformation at an
increasing rate with time, under constant deviatoric stress conditions is indicative of the
onset to failure.

7.9 “ABNORMAL” D EFORMATION BEHAVIOUR DURING CONSTRUCTION


OF EARTH AND EARTH-ROCKFILL EMBANKMENTS
From Section 7.4, the embankments highlighted as outliers to the general deformation
trend of the core showed up in both the plots of vertical strain profile and of total
settlement. All cases are from zoned embankments and mostly of thin to medium core
widths (combined core width less than 1H to 1V) with the core placed close to or wet of
Standard optimum moisture content.
The outliers to total core settlement at the end of construction (Figure 7.23 of
Section 7.4.3.3) include Hirakud, Beliche, Blowering, Nurek and Tedorigawa dams.
For Hirakud, Beliche, Blowering and Tedorigawa dams the difference in total
settlement from the trendline representative of the core type and width (Table 7.14) was
3 to 10 times the standard error and clearly very much greater than the estimated mean
for “normal” behaviour. For Nurek dam the difference above the mean was 2.7 times
Chapter 7: Deformation behaviour of embankment dams Page 7.188

the standard error, however, at a height of 290 m Nurek is the highest dam in the data
set and therefore the settlement estimate based on the trendline is not accurate.
The outliers to vertical strain profile (Figure 7.18, Figure 7.21 and Figure 7.22)
include Hirakud, Beliche, Blowering, Tedorigawa, Chicoasen and possibly Maroon
dams. For these embankments the vertical strain at end of construction for portions of
the core zone are greater than for the majority of case studies taking into consideration
core material type, moisture content at placement and core width.

7.9.1 PLASTIC DEFORMATION OF THE CORE DURING CONSTRUCTION

For Beliche, Tedorigawa, Chicoasen and Maroon dams large plastic deformation of the
wet placed core is considered to be the main reason for high vertical strains and/or high
total settlement at end of construction. At Nurek and Blowering dams plastic
deformation of the wet placed core is also considered to be a significant factor, but other
factors are also potentially significant. Beliche, Tedorigawa, Chicoasen and Nurek
dams are all zoned earth and rockfill dams with thin cores of clayey gravels placed at
about Standard optimum or wetter.
At Beliche dam (Figure 7.103) high vertical strains at end of construction, in the
range 5 to 7%, were measured for a large portion of the core from 16 m depth below
crest to foundation level at 55 m depth (Figure 7.21a). The vertical strains are well in
excess of those measured in similar thin, wet placed clayey sand to clayey gravel cores,
as well as those within thin, wet placed clay cores (Figure 7.22a) over a similar depth
range. Large lateral displacements of the core, as indicated by distortion of the
inclinometer tubes up and downstream of the core (Naylor et al 1997), are considered to
be a significant factor in large vertical strains measured. The high compressibility of
the lightly compacted, highly weathered schists and greywackes used as the inner
rockfill zone is likely to have been a contributing factor to the large lateral
displacements of the thin core. Another contributing factor to the potentially large
lateral displacement could have been collapse compression of the upstream rockfill on
saturation following partial impoundment of the reservoir to elevation 29 m (i.e. above
the crest level of the upstream cofferdam) following heavy rainfalls in January 1985
prior to completion of construction. However, no significant acceleration was recorded
in the vertical strain rate for the cross-arms represented in Figure 7.20a. Further details
of the deformation behaviour at Beliche dam are given in Section 1.3 of Appendix G.
Chapter 7: Deformation behaviour of embankment dams Page 7.189

Figure 7.103: Cross section through Beliche dam at the main section (Maranha das
Neves et al 1994)

Figure 7.104: Cross section of Chicoasen dam at the main section (Moreno and Alberro
1982)

At Chicoasen dam (Figure 7.104), a 260 m high central core earth and rockfill dam
constructed in a broad gully with near vertical abutment slopes, high vertical strains
within the core at IVM I-B4 of 5.5 to 6.5% were recorded over the depth range 85 to
105 m depth below crest level (Figure 7.21a). This region of the core is located within
the plastic region identified by Moreno and Alberro (1982). In comparison to Beliche
and Tedorigawa dams the vertical strains at Chicoasen dam are not overly high and are
isolated to a small region of the core. This is possibly reflective of smaller lateral
displacements of the core due to the likely low compressibility of the well-compacted,
moderately wide to wide gravelly filter / transition zones and well-compacted inner
rockfill shoulders. It is notable that the region of high vertical strain in IVM I-B4 at 85
to 105 m depth below crest level (elevation 295 to 315 m) is coincident with the base of
Chapter 7: Deformation behaviour of embankment dams Page 7.190

the core region where the filter / transition width is relatively thin (Figure 7.104).
Further details are given in Section 1.4 of Appendix G.
For Tedorigawa dam (Kawashima and Kanazawa 1982), a 148 m high zoned earth
and rockfill dam, limited information was available from the cited references on the
properties and placement methods of earth and rockfill materials. The high vertical
strains at end of construction (5 to 8%) over a large section of the core (from 50 to 125
m depth below crest level) as shown in Figure 7.21a are considered indicative of plastic
deformation of the core.
At Maroon dam, a central core earth and rockfill dam of 52 m height with thick core
of medium plasticity sandy clays to clayey sands placed on the wet side of Standard
optimum, relatively high vertical strains of 4.5 to 6.5% were measured in the depth
range 26 to 33 m below crest level at end of construction (Figure 7.22b) in all three
IVMs (ES1, 2 and 3). Given it is one of very few case studies in the database of zoned
embankments with wet placed thick clay cores (combined core width greater than 1H to
1V) and is only marginally outside a nominal limit established from five zoned
embankments with wet placed clay cores, it is difficult to really classify the vertical
strain as “abnormally” high.
At Blowering dam (Figure 7.124), a 112 m high central core earth and rockfill dam
with medium sized core of sandy clays placed on average from 0.2% dry to 0.3% wet of
Standard optimum, vertical strains in the core at end of construction ranged from 6 to
12% (Figure 7.22b) in the depth range 32 to 76 m below crest level. This region of the
core coincides with core placement on the wet side of Standard optimum. The lateral
displacement ratio at 73 m depth below the crest (Figure 7.12) was estimated at 2.3% at
end of construction. At the corresponding depth the vertical strain at end of
construction was almost 7% indicating the lateral displacement contributed to about
33% of the measured vertical strain at this depth. The relatively low modulus of the
supporting rockfill (estimated from HSG records in the rockfill) is thought to be
significant in the relatively high vertical strains measured in the core. Blowering dam is
discussed further in Section 7.10.3 and Section 6 of Appendix G.
In summary, embankments with high vertical strains within the core where plastic
deformations associated with lateral spreading are considered to be significant
contributing factor, have the following properties:
• Cores generally comprise clays, clayey sands and clayey gravels placed at moisture
contents at or wet of Standard optimum.
Chapter 7: Deformation behaviour of embankment dams Page 7.191

• The region of high vertical strain is generally observed over a large portion of the
core (e.g. Beliche, Tedorigawa and Blowering). Changes in the embankment zoning
geometry or moisture placement conditions of the core may restrict the region of the
core where high vertical strains are measured.
• Zoned embankments with relatively compressible rockfill shoulders, such as
observed at Beliche and Blowering dams.

7.9.2 COLLAPSE COMPRESSION OF THE CENTRAL EARTHFILL ZONE

The vertical deformation behaviour of the core of Hirakud dam stands out as an outlier
in terms of total settlement (Figure 7.23) and vertical strain in the mid to lower portion
of the core (Figure 7.18b).
The very broad central earthfill zone at Hirakud dam (Figure 7.105) comprised
mostly low to medium plasticity clayey gravels to clayey sands placed 1% to 3% dry of
Standard optimum. Vertical strains in the central earthfill zone at IVM C at end of
construction, installed in the deeper gully section of the embankment, reached values in
the range 7.5% to 13.5% in the lower 15 m of the core (35 to 49 m) as shown in Figure
7.18b. These values are well in excess of the typical vertical strains recorded at end of
construction for similar dry placed earthfills.
During the shutdown period of the 1952 monsoon season high water levels were
impounded in the reservoir and backwaters engulfed the embankment section within the
gully region (about 12 to 13 m height at this time) from July to December 1952 (Rao
1957). The large settlement during the 1952 shutdown period (Figure 7.106) is likely
due to collapse settlement of the dry placed earthfill on saturation during inundation.
Subsequent large settlements during the following two construction periods in 1953 and
1954 were predominantly within the lower 15 m of the core and likely due to the low
moduli of the wetted earthfill. At end of construction the total settlement was in the
order of 2170 mm, 1430 mm (or 66%) of which occurred within the lower 15 m or
lower 30% of the embankment. Investigations at end of construction (Rao 1957)
encountered “soft soil patches” in the lower 21 m where moisture contents were as high
as 10% above Standard optimum. Further details are given in Section 3.2 of Appendix
G.
Chapter 7: Deformation behaviour of embankment dams Page 7.192

Figure 7.105: Section at Hirakud dam (Rao and Wadhwa 1958)

0 60
Cumulative Settlement (mm)

Approx. Fill Elevation (m) .


500 48

1000 cumulative settlement at IVM C 36


embankment elevation
period of inundation
.

1500 during 1952 monsson 24


season

2000 12

2500 0
0 1 2 3 4
Time (years since Jan 1952)
Figure 7.106: Cumulative settlement during construction at Hirakud dam (adapted from
Rao 1957)

7.9.3 RESERVOIR FILLING DURING CONSTRUCTION

Partial impoundment during construction can affect the vertical strains and total
settlement of the core for zoned embankments with free draining upstream shoulders
that are susceptible to collapse compression on wetting. As the monitoring records at
several dams (Section 7.10.2) and results of the finite difference modelling (Section
7.5.2) indicate, embankments with wet placed cohesive earthfill cores of low undrained
shear strength are susceptible to high vertical strains in the core resulting from collapse
compression of the upstream rockfill on first filling.
At Nurek dam (Figure 7.107), a 290 m high zoned earth and rockfill dam with thin
central core zone, the total settlement of the core during construction is considered to
have been affected by the reservoir impoundment during construction. The thin, sandy
clayey gravel core was likely to have been placed close to or wet of Standard optimum
given the high pore water pressures developed (Sokolov et al 1985). Reservoir filling
Chapter 7: Deformation behaviour of embankment dams Page 7.193

was undertaken at construction proceeded. Sokolov et al (1985) report total settlements


at end of construction of 13.7 m for the core, 11.9 m for the upstream shoulder and 4 to
6.5 m for the downstream shoulder. They comment that collapse compression on
wetting in the gravel to bouldery upstream shoulder fill contributed to its large
settlement relative to the downstream shoulder.
The large settlement of the thin, wet placed sandy clayey gravel core is likely to
have been affected by the collapse compression within the upstream shoulder. As
discussed in Section 7.5.2 (Figure 7.35), collapse compression in the upstream shoulder
can result in plastic deformation from lateral spreading in cores of low undrained shear
strength with subsequent large vertical strains in the core.

Figure 7.107: Cross section of Nurek dam (adapted from Borovoi et al 1982)

7.9.4 SHEAR SURFACE DEVELOPMENT IN THE CORE DURING CONSTRUCTION

High localised vertical strains in the core can be an indicator of possible development of
a shear surface. The clearest example of this in the literature is probably the failure at
Carsington dam (Skempton and Vaughan 1993; Rowe 1991; Potts et al 1990).
Carsington dam, a zoned earthfill embankment of 36 m maximum height, failed
during construction in early June of 1984. The central core, including its unusual
“boot” structure on the upstream side, comprised high plasticity clays that were placed
well wet of Standard optimum moisture content and heavily rolled. The outer earthfill
zones were of weathered mudstone placed in thin layers and well compacted.
The deformation records from IVM gauges installed in the central core highlighted
the development of the shear surface in the core prior to the failure. At IVM C, located
4 m upstream of the axis at chainage 850 m, the vertical strain profile (Figure 7.109a)
Chapter 7: Deformation behaviour of embankment dams Page 7.194

shows an increase from 7 to 9% at about elevation 180 m from early to late October
1983 during the early stages of the winter shutdown. After this date, and during the
shutdown period, the tube at IVM C became constricted and then blocked at this
elevation indicating the continuance of shear type deformation. This zone of high
vertical strain, at about elevation 180 m, is coincident with the surface of rupture
through the clay core (Figure 7.108) determined from investigation after the failure.
At IVM B, located 4 m upstream of the axis at chainage 705 m, the vertical strain
profile (Figure 7.109b) shows a large increase in vertical strain between elevations 185
and 192 m in the days immediately prior to and during the development of the failure.
Figure 7.110 (from Rowe 1991) highlights the large shear strain development in
IVMs B and C leading up to the failure. This type of plot of vertical strain versus bank
level or height above the gauge is useful in identification of potential shear development
as it highlights the development of large localised strains that are not necessarily
attributable to increasing total vertical stress or to consolidation type settlements. This
behaviour contrasts that for a number of the case studies from the database (Figure
7.20), which typically show a constant or decreasing rate of vertical strain with
increasing fill height above the gauge.
Further details on the deformation behaviour from IVM gauges leading up to the
failure at Carsington dam are given in Section 3.1 of Appendix G.
At Blowering dam the vertical strain deformation behaviour in the core during
construction possibly also shows the development of a shear surface. The deformation
behaviour of Blowering dam is discussed in Section 7.10.3.1 and Section 6 of Appendix
G.

Figure 7.108: Carsington dam, section at chainage 825 m after failure (adapted from
Skempton and Vaughan 1993).
Chapter 7: Deformation behaviour of embankment dams Page 7.195

a) 210 b) 210
IVM C at Chainage 850 m IVM B at Chainage 705 m

200 200

Early Oct. 1983


190 26 Oct. 1983 190
Elevation (m)

Elevation (m)
180 180
foundation level

170 170 Early Oct. 1983


24-May-84
4-Jun-84
foundation level

160 160
0% 3% 6% 9% 12% 0% 3% 6% 9% 12%
Vertical Strain (%) Vertical Strain (%)

Figure 7.109: Carsington dam, vertical strain profiles from IVM during construction, (a)
IVM C at chainage 850 m, and (b) IVM B at chainage 705 m (adapted from Rowe
1991).

Figure 7.110: Carsington dam, variation of the maximum vertical strain in the core with
bank level (Rowe 1991).
Chapter 7: Deformation behaviour of embankment dams Page 7.196

7.10 “ABNORMAL” D EFORMATION BEHAVIOUR P OST CONSTRUCTION


OF ZONED EARTH AND ROCKFILL EMBANKMENTS
In Sections 7.5.3 and 7.6 aspects of the post construction deformation behaviour of a
large number of case studies of zoned earth and rockfill dams, mainly central core earth
and rockfill dams, were labelled as “abnormal” or possibly “abnormal”. Of these, the
deformation at several embankments was clearly an outlier to the general trend
including Ataturk, Beliche, Svartevann, Eppalock and Djatiluhur dams. At Eppalock
and Djatiluhur dams consultants to the owners (Woodward Clyde 1999; Sowers et al
1993) indicated that the upstream slope under a drawn down reservoir condition was
approaching a marginal stability condition.
The other case studies where one or more aspects of the deformation behaviour were
labelled as possibly “abnormal” included Gepatsch, South Holston, Glenbawn,
Srinagarind, Canales, El Infiernillo, Eildon, Upper Yarra, Matahina, Blowering,
Wyangala, Bellfield and Cougar dams. Others named for unusual internal deformation
behaviour in the core include Copeton and La Grande 2 dams. In total this is
approximately 20 of 75 zoned earth and rockfill embankments with thin to thick core
widths, most of which are central core earth and rockfill dams. For virtually all but
Eppalock dam (which has been remediated with upstream fill to improve stability) and
Djatiluhur dam the overall stability of the embankment is not in question. The
potentially “abnormal” deformation trends include:
• Accelerations in settlement rate over short periods of time of SMPs on the crest and
sometimes the upstream slope, often, but not always, post first filling on drawdown.
This is an attribute of quite a number of the case studies, including Eppalock and
Djatiluhur dams and therefore potentially indicative of marginal stability conditions
(refer also to San Luis and Belle Fourche dams discussed in Section 7.11.2).
• Non-recoverable displacements over short time periods post first filling (usually
upstream on drawdown), a change in the direction of the displacement trend, and
high rates of displacement long-term. Once again some of these trends are evident
in the case studies approaching marginal stability.
• Large magnitudes of settlement or displacement compared to similar types of
embankments. The direction of displacement is also a factor.
• High long-term settlement rates.
Chapter 7: Deformation behaviour of embankment dams Page 7.197

Of the case studies named, a large number incorporate rockfills that are susceptible
to large deformations due to collapse compression on wetting and/or the rockfill is of
high compressibility. These include rockfills that are poorly compacted, dry placed and
poorly to reasonably compacted, dumped and sluiced rockfills, weathered rockfills, and
rockfills of rock type susceptible to large loss in unconfined compressive strength on
wetting. The implication is that the rockfills susceptible to large deformations after the
end of construction can and do have a significant influence on the post construction
deformation behaviour of the overall embankment. Conversely, the data indicates that,
in most cases, sound rockfills that are wetted and well and reasonably to well
compacted are not susceptible to large collapse compression on first filling and the
overall post construction deformation behaviour of the embankment is “normal” and
generally of limited magnitude.

7.10.1 ROCKFILL SUSCEPTIBILITY TO COLLAPSE COMPRESSION

Most rockfills in the upstream shoulder undergo some collapse compression on wetting
during first filling. In most cases the amount of collapse compression and its effect on
the overall deformation behaviour of the embankment is limited to negligible. As
previously discussed (Section 7.5.2), the factors affecting the susceptibility of a rockfill
to collapse compression include the method of placement (layer thickness, roller type,
number of passes), moisture content at placement, effective stress level, particle size
distribution, particle shape (angular versus rounded), degree of weathering of the rock,
and the rock type itself (its loss in unconfined compressive strength on wetting).
Table 7.25 presents a list of embankments for which collapse compression of the
rockfill resulted in moderate to large settlements of the upstream shoulder on first
filling. From the data, and other information from the published literature, it is evident
that:
• Dry dumped rockfills are susceptible to very large settlements due to collapse
compression when wetted. At Cogswell dam the dry dumped rockfill settled up to
6% due to wetting from heavy rainfall and later sluicing (Baumann 1958).
• Dry dumped and poorly sluiced rockfills are also susceptible to large settlements
when wetted. Howson (1939) describes the large settlements that occurred at
Strawberry and Dix River dams on partial flooding of the dumped and poorly
sluiced rockfill. At Dix River dam the settlement amounted to 0.2 to 0.3% of the
Chapter 7: Deformation behaviour of embankment dams Page 7.198

height of rockfill, but was more likely closer to 1% in the flooded portion of the
rockfill.
• Dry placed and poorly to reasonably compacted rockfills are susceptible to very
large settleme nts due to collapse compression when wetted. Settlements of the
upstream shoulder at Eppalock, Eildon and Gepatsch dams, and the downstream
shoulder at Svartevann dam were all close to or greater than 1% on first filling. At
Eildon and Eppalock dams settlements much greater than the measured 1.23% and
0.90% respectively were likely to have occurred because the monitoring missed
about the first half of reservoir filling. Internal vertical strains in the rockfill at
Svartevann dam were as much as 1.4 to 2.1% at four years after construction.
• Dumped and sluiced rockfills are susceptible to large settlements due to collapse
compression when wetted. At Cherry Valley and Mud Mountain dams measured
settlements of the upstream slope were close to 1%. At Watuaga, Nottely and South
Holston large settlements of the upstream rockfill also occurred on first filling.
• Some weathered compacted rockfills are susceptible to very large settlements due to
collapse compression when wetted, presumably if they are placed with limited
quantities of water. At Beliche dam, vertical strains of up to 2.1% were measured
on first filling within the “lightly compacted” rockfill of weathered schists and
greywackes. At Ataturk dam the settlement of the upstream weathered rockfill was
potentially very large as indicated by the very large post construction settlement of
the crest (4% at almost 7 years after construction).

The use of watering during placement of rockfill reduces the susceptibility of the
rockfill to collapse compression as implied in the above summary. Rock type does not
appear to stand out as a significant factor, being over-shadowed by placement method
and variability in the data. However, the large settlements attributed to collapse
compression in the weathered rockfills at Ataturk and Beliche dams tends to indicate
that reduction in the rock substance strength due to wetting is a significant influence on
susceptibility to collapse compression.
Chapter 7: Deformation behaviour of embankment dams Page 7.199

Table 7.25: Embankments for which collapse compression caused moderate to large
settlements.
Settlement Maximum Rockfill - Upstream Shoulder
Height
on First Internal n
Dam Name Comp Water at Comment
Filling Strain Type
(m) 1 2 Rating *3 Placement
(%) * (%) *
weathered basalt & suspect very high because of very
Ataturk 184 very high reas (?) 2 to 6%
limestone large settlement of the crest

weathered schists very high strains in weathered


Beliche 55 1.15 2.1 poor unknown
& greywacke rockfill
4 well (3A) monitoring missed first 50 m of
Blowering 112 > 0.32 * quartzite & phyllite yes
reas (3B) filling
settlement of upstream edge of
Canales 156 0.64 limestone unknown unknown
crest
granite & poor
Cherry Valley 100 1.05 sluiced
granodiorite (dumped)
reas to well greater because cofferdam was
4
Copeton 113 > 0.5 * granite (3B) dry overtopped prior to end of
poor (3C) construction
greater because reservoir within
4
Eildon 79 > 1.23 * quartzitic sandstone poor dry ? 33 m of FSL before monitoring
started.
diorite & silicified lower 50 to 75m saturated prior to
El Infiernillo 148 0.38 0.7 poor to reas dry
conglomerate end of construction.
4 monitoring missed filling of first 20
Eppalock 47 > 0.90 * basalt poor dry
m.
some influence of the alluvial
Gepatsch 153 0.8 to 1.0 gneiss reas dry
foundation
Glenbawn large collapse settlement on first
76.5 0.83 limestone poor dry
(main dam) filling
poor
Mud Mountain 128 1.05 andesite & tuff sluiced
(dumped)

1.0 1.4 to 2.1 settlement in downstream shoulder


Svartevann 129 granitic gneiss reas dry
(downstream) (downstream) in first 4 years after construction.

4 quartz, granite & suspect large, missed filling to


Vatnedalsvatn 121 > 0.37 * reas dry
gneiss within 10 m of FSL

Notes: *1 settlement of upstream shoulder on first filling unless stated


*2 internal vertical strains measured in the upstream rockfill during first filling unless stated
*3 Compn Rating = compaction rating of rockfill; “well” = well-compacted, “reas to well” =
reasonably to well compacted, “reas” = reasonable compaction, “poor” = poorly compacted.
Refer Section 1.3.3 for definitions of the terms.
*4 “>” indicates settlement likely to be greater than that stated because part of part of the
settlement on first filling was not measured.

For well and reasonably to well compacted rockfills collapse compression, in most
cases, has a limited to negligible influence on the settlement of the upstream shoulder
on first filling, particularly for rockfills sourced from sound rock types and watered
during placement. But, for several embankments collapse compression was considered
to have had some influence on the settlement of the upstream shoulder on first filling, as
indicated by either the relatively large magnitude of settlement (greater than about 0.3%
on first filling) and/or a small but significant differential settlement between the up and
downstream shoulder (great than about 0.1% difference to upstream). Examples from
the database include:
Chapter 7: Deformation behaviour of embankment dams Page 7.200

• Several dams with dry placed (and well or reasonably to well compacted) rockfill
sourced from sound rock types, including LG-2, Round Butte and possibly
Dartmouth and Thomson dams. These embankments are all greater than 130 m in
height. Rock types varied from granitic gneiss to basalt to sedimentary sandstones
and siltstones.
• Several dams with lesser quality rockfills (including rockfills sourced from
weathered rocks or rock types of medium to high unconfined compressive strength,
or rockfills with high fines content), including:
− The 146 m high La Angostura zoned earth and rockfill dam. Much greater
settlements were measured for the upstream shoulder on first filling (0.38%
compared to 0.10 to 0.18% for the downstream shoulder). The embankment
zoning geometry may have partly contributed to the greater settlement, but
collapse compression in the well compacted but poor quality (as described by
Benassini et al (1976)) limestone rockfill is likely to be the main reason.
Benassini et al (1976) indicate the rock used as rockfill was highly contaminated
and susceptible to particle breakage on trafficking, and the rockfill as compacted
was of high compressibility and low shear resistance.
− The 90 m high Dalesice dam where much greater settlements were measured for
the upstream shoulder than the downstream shoulder (0.39% compared to 0.12
to 0.18%) in the first 15 years after construction. Brousek (1976) indicates that
on opening the quarry a considerable proportion of the rock material was of
worse quality that originally presupposed, and might have been a factor in the
greater settlement of the reasonably to well compacted upstream rockfill
shoulder. The rock type and whether or not water was added during placement
are not known.
− The 112 m high Blowering dam where greater settlements on first filling were
measured for the upstream than downstream shoulder (0.32% compared to
0.13%). Settlement of the upstream shoulder is likely to have been greater as
post construction monitoring did not commence until the reservoir was more
than 50% filled. Collapse compression within the reasonably compacted outer
Zone 3B and possibly well compacted inner Zone 3A rockfill due to large loss in
rock strength on saturation was considered to contribute to the greater settlement
of the upstream shoulder, even though large volumes of water were used during
placement. The source rock, particularly the phyllite rock used in the Zone 3B
Chapter 7: Deformation behaviour of embankment dams Page 7.201

rockfill, was susceptible to large strength loss on wetting, up to 60% of its dry
strength.
• At the 35 m high Glenbawn saddle dam (Saddle Dam A) relatively large settlement
of both shoulders occurred during first filling, but the settlement of the upstream
shoulder was greater (0.40% compared to 0.28%). Possible reasons for the likely
collapse compression are not known as only limited details were gathered on the
material type and placement methods of the rockfill.

7.10.2 DEFORMATION ON FIRST FILLING IN EMBANKMENTS WHERE COLLAPSE


COMPRESSION OCCURS IN THE UPSTREAM ROCKFILL SHOULDER

The influence of collapse compression on wetting of the upstream rockfill shoulder


during first filling on the deformation behaviour within zoned earth and rockfill
embankments was discussed in Section 7.5.2. Two bounds of behaviour were defined:
(i) Embankments where the core is of high undrained strength and relatively low
drained compressibility, such as a partially saturated clayey core of very stiff to
hard strength consistency or well compacted sandy to gravelly core. On first
filling the upstream rockfill settles relative to the core and the down drag due to
differential settlement results in the development of high shear stresses at the
upstream core / shoulder interface as described by Squier (1970).
(ii) Embankments with wet placed clayey cores of low undrained strength. On
collapse compression in the upstream shoulder the core settles with the upstream
shoulder and large differential settlements occur at the downstream core / shoulder
interface (Figure 7.35). Deformations within the core are largely plastic.

The actual deformation behaviour is much more complex than the idealised
simplified models, but these two bounds of behaviour are observed in a number of the
case studies where collapse compression of the upstream rockfill is significant during
first filling. Examples from the database are discussed below.
An important aspect of the deformation behaviour on first filling shown in the
monitoring records at the well instrumented El Infiernillo dam (refer Section 7.10.2.1
below) and Chicoasen dam (refer Section 1.4 of Appendix G) is the deformation
behaviour of the upstream filter zones when collapse compression occurs within the
upstream rockfill. For both embankments localised regions of high vertical strain
Chapter 7: Deformation behaviour of embankment dams Page 7.202

developed within the upstream filters indicating the formation of shear surfaces. The
mechanism for development of the shear surfaces is considered to be due to the high
shear stresses that develop within the filter as a result of the differential settlement
associated with collapse compression of the upstream rockfill shoulder and shedding of
load onto the relatively high modulus filters.

7.10.2.1 Collapse Compression on First Filling and its Influence on the Deformation
Behaviour of Wet Placed Clayey Cores.

The following discussion presents a summary of case study evidence of wet placed
clayey cores where significant collapse compression occurred within the upstream
rockfill shoulder on wetting and the core largely deformed with the upstream shoulder.
In several of the cases there is evidence that differential settlements were concentrated
at the downstream interface of the core, and at others the internal deformation records
indicate large differences in settlement between the core and downstream rockfill
shoulder suggesting likely concentration at the downstream interface. The case studies
discussed below include El Infiernillo, Djatiluhur, Canales and Beliche dams. A similar
mechanism is considered to have occurred at La Angostura and Netzahuacoyotl dams,
but data records found in the published literature were limited.
The deformation behaviour at Chicoasen dam, also discussed below, indicates that
this behaviour can occur where settlements due to collapse compression of the upstream
rockfill are relatively small.

(a) Deformation Behaviour at El Infiernillo Dam During First Filling.


Marsal and Ramirez de Arellano (1967), Squier (1970) and Nobari and Duncan (1972b)
discuss the deformation behaviour on first filling at El Infiernillo dam. In the following
summary only the vertical deformation behaviour measured from internal deformation
gauges and the interaction between various zones in the embankment is discussed.
Further details are presented in Section 1.9 of Appendix G.
El Infiernillo dam (Figure 7.111), constructed in the early 1960’s, is a central core
earth and rockfill dam of 148 m maximum height located in a narrow valley with steep
abutment slopes. Well-compacted filter / transition zones are located either side of the
wet placed, high plasticity sandy clay core. Rockfill, of quarried diorite and silicified
Chapter 7: Deformation behaviour of embankment dams Page 7.203

conglomerate, was placed dry and track rolled by bulldozer; Zone 3A in 0.6 to 1.0 m
lifts and Zone 3B in 2.0 to 2.5 m lifts.
The section of the upstream rockfill zone between the embankment and the
upstream cofferdam was flooded during construction to reduce the impact of collapse
compression on first filling.

Figure 7.111: Maximum section at El Infiernillo dam (Marsal and Ramirez de Arellano
1967).

The measured internal settlements of the rockfill, filters and core at Station 0+135 at
selected time intervals over the period of first filling are presented in Figure 7.113
(instrument locations are shown in Figure 7.112). The plots highlight several important
aspects of the deformation behaviour:
• The settlement profiles of the core, upstream filter and upstream rockfill are similar,
and are different to that of the downstream rockfill.
• At end of first filling, large and uniform vertical strains were measured in the
upstream rockfill between elevation 80 and 125 to 130 m. Similarly, large vertical
strains were measured in the core, but at a higher elevation, from 105 to 150 m.
• In the upstream filter (IVM I4) localised zones of high vertical strain were measured
at about elevation 102 m and 132 m.
• Low vertical strains were measured below about elevation 80 m in the upstream
rockfill. This is likely to be because due to the pre-saturation during construction
and therefore this region is not susceptible to collapse settlement post construction.

The settlement of the core and upstream filter / transition on first filling is largely
controlled by the collapse compression in the upstream rockfill. Differential settlements
Chapter 7: Deformation behaviour of embankment dams Page 7.204

between the upstream rockfill and filters will occur above about elevation 80 m and, as
a result of the down-drag effect, high shear stresses are developed on the upstream side
of the upstream filter zone. The deformation response within the upstream filter /
transition zone to these high shear stresses is the development of shear surfaces, as
evidenced by the localised regions of high vertical strain within gauge D1 at elevations
102 m and 132 m.
The vertical strain in the core between elevation 105 and 150 m is almost three
times the strain below elevation 105 m (average of 0.66% compared to 0.24% in
February 1966). It is considered that the greater vertical strain in the mid region of the
core is largely due to plastic type deformations in undrained loading as the core deforms
with the upstream shoulder. The strains in the core are not localised as they are in the
upstream filter because of the low undrained strength and plastic nature of the wet
placed clay core.
The vertical strain profile in the downstream rockfill contrasts that of the core,
upstream filter and upstream rockfill. Differential settlements between the core and
downstream filter above elevation 105 m are large at more than 200 mm, and indicate a
likely concentration at the downstream interface of the core.

Figure 7.112: El Infiernillo dam, deformation instrumentation on station 0+135 at the


lower left abutment (Marsal and Ramirez de Arellano 1967).
Chapter 7: Deformation behaviour of embankment dams Page 7.205

I1 I4
a) 180 b) 180 I1 I4
15 June 1964 15 June 1964
(0.57 yrs post) (0.57 yrs post)
D1
D2 D1
D2
160 160

140 140
Elevation (m)

Elevation (m)
120 120

100 100

9 July 1964
80 80
(0.50 years post)
18 Dec 1964
(1.01 years post)
60 60
0 100 200 300 400 500 0 100 200 300 400 500
Settlement (mm) Settlement (mm)

c) 180 I1 I4
15 June 1964
(0.57 yrs post) D1
D2
160

I4 - localised
zones of high
Legend
vertical strain
140 I1 - avg. 0.66% reservoir level
vertical strain
Elevation (m)

I inclinometer
(EL 105 to 143)
D internal settlement gauge
120 I1 centre of core (refer to
I1 I4 upstream filter
I4 D1 - 0.70% avg. vert. Figure 7.112
strain (EL 80 to 130) D1 upstream rockfill for locations)
100
D2 downstream rockfill

80
2 Feb 1966
(2.14 years post)
60
0 100 200 300 400 500
Settlement (mm)
Figure 7.113: El Infiernillo dam, internal settlements during first filling at Station
0+135.

(b) Deformation Behaviour at Chicoasen Dam During First Filling (Section 1.4 of
Appendix G).
Chicoasen dam (Figure 7.114), constructed in the late 1970’s, is a central core earth and
rockfill dam of 261 m maximum height located in a narrow valley with near vertical
abutment slopes. The core comprises well-compacted clayey gravels placed at close to
Standard optimum moisture content. Filter / transition zones of well-compacted gravels
Chapter 7: Deformation behaviour of embankment dams Page 7.206

are located either side of the narrow core and the main rockfill zone (Zone 3A) consists
of well-compacted quarried limestone.
The embankment was well instrumented at the main section (Figure 7.115) with
numerous inclinometers and cross-arms installed in the core, filters and rockfill zones.
Moreno and Alberro (1982) present a selection of the instrumented deformation
behaviour on first filling and a summary of the vertical deformation is presented below.

Figure 7.114: Main section at Chicoasen dam (Moreno and Alberro 1982).

Figure 7.115: Chicoasen dam, inclinometer and cross-arm locations at the main section
(Moreno and Alberro 1982).

The internal vertical settlement profiles for the period of first filling are presented
Figure 7.116. Note that the dates of the monitoring period are different between gauges,
but cover most of the period of first filling from 1 May 1980 to late July 1980. As
shown, a region of high vertical strain developed in the upstream rockfill between
elevations 260 m and 300 m (strains of 0.4 to 0.5% in IVM I-A5 and I-A6), possibly
Chapter 7: Deformation behaviour of embankment dams Page 7.207

due to collapse compression on wetting in the well-compacted limestone rockfill,


although Moreno and Alberro (1982) comment that it may be due to plastic type
behaviour on reduction in effective horizontal stresses.
Vertical strains in the upstream filter zones are much greater above elevation 270 m
and are most likely in response to the higher stresses developed within the filter as a
result of differential settlement following collapse compression in the upstream rockfill.
In the Zone 2A filter (IVM D5) localised concentrated zones of vertical deformation
were evident at elevations 285 m and 317 m, and are likely to be due to shear type
deformations. The vertical settlement profile in the core shows that vertical strains are
much greater above about elevation 290 m (0.5% to 0.75% compared to 0.3% below
elevation 290 m), which is above the base elevation of high strains in the upstream
rockfill and upstream filters. The broad zone over which high vertical strains developed
in the core suggests it is largely due to plastic type deformations.
The deformation behaviour suggests that the collapse type deformation of the
upstream rockfill has a controlling influence on the deformation in the upstream filter /
transition zones and core.

400
March to May 1980

350
Elevation (m)

300

250

200 I-A5, upstream rockfill (14/4/80 to 16/6/80)


I-A6, upstream rockfill and filters (22/3/80 to 8/7/80)
I-AB, centre of crest (30/3/80 to 2/01/81)
D2, upstream Zone 2B filter (15/5/80 to 21/1/81)
D5, upstream Zone 2A filter (14/5/80 to 21/1/81)
150
0 100 200 300 400 500 600 700 800 900
Settlement (mm)
Figure 7.116: Internal settlement profiles during first filling at Chicoasen dam (adapted
from Moreno and Alberro 1982).
Chapter 7: Deformation behaviour of embankment dams Page 7.208

Moreno and Alberro (1982) comment that no concentration of deformation was


measured in the downstream rockfill and it is assumed that the deformations would have
been relatively small. Given the high vertical strains in the core it is likely that
differential settlements were concentrated between the core and downstream shoulder,
but Moreno and Alberro (1982) give no indication that cracking or differential vertical
displacements were evident at the crest.

(c) Deformation During First Filling at Beliche Dam (Section 1.3 of Appendix G)
Beliche dam (Figure 7.117), Portugal is a central core earth and rockfill embankment of
55 m maximum height that was completed in 1986. The compacted earthfill core of
clayey sandy gravels was placed at moisture contents close to Standard optimum
moisture content. Naylor et al (1997) indicate that the inner rockfill zone (Zone 3A) of
highly weathered and fractured schists and greywackes was placed in 1.0 m layers,
“relatively lightly compacted” and comprised a “significant proportion” of fines. The
outer rockfill zone (Zone 3B) was of “good quality” greywackes placed in 1.0 m layers
and also “relatively lightly compacted”. Water is indicated as being added to the
rockfill, but in what proportion or to which zones is not clear.
During construction the reservoir level exceeded the height of the upstream
cofferdam and saturated the upstream rockfill to elevation 29 m.
The internal vertical settlement profiles of the core and up and downstream rockfill
shoulders during the period of first filling after construction are shown in Figure 7.118.
Very large vertical strains (average of 2.1%) were measured on first filling above
elevation 27 m in the inner upstream rockfill zone of poorly compacted weathered rock,
largely due to collapse compression on wetting. Large vertical strains were not
measured below about elevation 27 m because the upstream rockfill was saturated to
this elevation during construction. The vertical settlement profile in the core is similar
to that of the inner upstream rockfill shoulder. Relatively low vertical strains (average
0.7%) were measured below about elevation 29 m and very large vertical strains above
elevation 29 m, particularly in the upper 10 to 12 m of the core where they averaged
3.2%. The deformation of the wet placed clayey gravel core is considered to be largely
due to undrained plastic type deformation, and to be largely controlled by collapse
compression in the upstream rockfill.
Relatively large settlements also occurred in the downstream rockfill shoulder over
the monitored period, possibly due to rainfall induced collapse compression.
Chapter 7: Deformation behaviour of embankment dams Page 7.209

Figure 7.117: Beliche Dam, cross section (Maranha das Neves et al 1994)

crest elevation

50

surface elevation at IVM I1


average of 3.2%
vertical strain

40 surface elevation at IVM I6


Elevation (m) .

30 average of 2.1%
vertical strain (I1)

20

IVM I1 - upstream rockfill


10
IVM I3 - central core
IVM I6 - downstream rockfill

0
0 100 200 300 400 500 600 700
Settlement (mm)
Figure 7.118: Beliche dam, post construction internal vertical settlement profile within
the embankment for the period from end of construction to 2.75 years post construction.

(d) Deformation During First Filling at Canales Dam (Section 2.1 of Appendix G)
The deformation behaviour on first filling at Canales dam is a clear example of the
development of differential settlement at the downstream interface of the core. Canales
dam (Figure 7.119) in southern Spain is a zoned earth and rockfill embankment
constructed in a narrow, steep sided valley. The embankment is of 156 m maximum
Chapter 7: Deformation behaviour of embankment dams Page 7.210

height and was constructed in two stages; the first stage of 100 m (to elevation 910 m)
from 1979 to 1981, and the second stage to crest level from 1985 to 1986.
The thin core (Bravo 1979) consisted of high plasticity silty clays placed at moisture
contents on the wet side of Standard optimum (from OMC to 2% wet of OMC). Very
broad transition zones of clayey to silty gravelly sand are located either side of the core
and the rockfill shoulders are of quarried limestone. Bravo (1979) indicates the
embankment materials were compacted to “the highest possible density”, but it is not
known what layer thickness was used or if water was added during placement.
During the period of first filling a substantial longitudinal crack developed in the
crest between the core and downstream transition (Figure 7.120a). Giron (1997)
indicates that the crack was first observed in 1989 (2 years after end of construction)
when the recorded differential settlement of the crest was about 200 mm (Figure
7.120b). On reservoir raising to elevation 930 m at 3.8 years post construction a vertical
slump had clearly developed across the downstream core / shoulder interface with a
differential settlement of 405 mm. On reservoir raising to full supply level (elevation
958 m) for the first time, the rate of settlement of the mid to upstream portion of the
crest increased significantly and differential settlement across the crack increased to
about 1000 mm. Giron (1997) and Bravo et al (1994) attribute the crest settlement
behaviour to collapse settlement of the upstream shoulder fill on saturation.

Figure 7.119: Typical section of Canales dam (Bravo 1979).


Chapter 7: Deformation behaviour of embankment dams Page 7.211

b) 0 960

200 935

Reservoir Level (elevation, m) .


d
Settlement (mm) .

400 910

600 885

800 860

1000 SMP B - centre of crest 835


SMP C - crest, downstream edge
Reservoir

1200 810
0 2 4 6 8 10 12
Time (years since end of construction)

Figure 7.120: Canales dam, (a) cracking and differential settlement at the crest, and (b)
post construction settlement at the crest (Giron 1997)

Several possibilities could explain the observed deformation behaviour:


• The low undrained strength of the wet placed, high plasticity clay core and
consequent plastic deformation as the core deforms with the upstream shoulder as it
settles due to collapse compression.
• A shear type movement in the core, along a defined plane of shearing with
backscarp at the downstream core / transition interface.
• A combination of the above.

The first explanation is considered more feasible than the second mainly because the
period of rapid crest settlement (after 9.6 years) occurs when the collapse settlement of
the upstream rockfill is localised to the upper 20 to 30 m of the upstream shoulder, and
when the reservoir level is close to full supply level where it would provide a high level
Chapter 7: Deformation behaviour of embankment dams Page 7.212

of support to the upstream face of the core. For this reason, shear type movements on a
pre-existing shear surface in the core seem less likely than plastic deformation of a wet
placed high plasticity clay core. However, a combination of both is considered possible.

(e) Deformation at Djatiluhur Dam (Section 1.7 of Appendix G)


Aspects of the deformation behaviour and observations during and after construction at
Djatiluhur dam in Indonesia indicate that the core largely deforms with the upstream
shoulder and large differential settlements occur at the downstream interface of the core.
Djatiluhur dam (Figure 7.121) is a central core earth and rockfill dam of 105 m
maximum height constructed in the early to mid 1960’s. The thin core was of high
plasticity clays derived from weathered claystone, placed at moisture contents on the
wet side of Standard optimum. Investigations after construction indicated the core to be
of low undrained strength, particularly above elevation 65 m (Sowers et al 1993).
Details on the placement methods of the rockfill, sourced from quarried andesite, are
not precisely known but are thought to include both roller compaction and placement
without formal compaction in layer thicknesses ranging from 0.5 m up to 2 m (Farhi and
Hamon 1967; Sherard 1973; Sowers et al 1993). Farhi and Hamon (1967) comment
that most of the rockfill in the mid to lower elevation was well sluiced (300% water by
volume), and in the upper section the rockfill was placed in 1 to 2 m lifts, 30% by
volume water added and trafficked by trucks and bulldozer.

Figure 7.121: Main section at Djatiluhur dam (Sowers et al 1993)

Because first filling largely occurred during the period of construction it is difficult to
gauge the magnitude of influence of collapse compression on the settlement of the
upstream rockfill. But, important aspects of the deformation behaviour were revealed
Chapter 7: Deformation behaviour of embankment dams Page 7.213

from monitoring during a shutdown period in construction and from test pits excavated
after construction as described by Sherard (1973).
In early January 1965 construction was halted when the embankment had reached
elevation 103 m. As described by Sherard (1973), shortly after construction was
stopped a longitudinal crack appeared at the boundary between the core and
downstream filter (Figure 7.122a), reaching a total length of some 500 m. Monitoring
points were then established at elevation 103 m and the measured deformation records
(Figure 7.122b) showed much larger settlements of the core compared to the rockfill
shoulders, and limited settlement of the downstream shoulder. Differential
displacements indicated lateral spreading of the core amounted to some 400 mm over
the period January to July 1965. On raising the embankment to design level in August
1965 the core settlement at elevation 103 m totalled some 800 mm, well in excess of
that measured on the downstream slope at a similar elevation, and relatively large
settlements were also recorded on the upstream slope at elevation 100 m of about 400
mm over this period.
The very large settlements of the core relative to the shoulders, particularly the
downstream shoulder, are considered to be largely due to undrained plastic type
deformations from lateral spreading of the core. The large differential settlements
between the core and both the upstream and downstream shoulder indicates that arching
or stress transfer from the core to the shoulders would have occurred. The low vertical
stresses within the core were confirmed by water pressure testing after construction
(Sherard 1973), which showed that under relatively low water pressures the horizontal
cracks present in the upper region of the core would opened up resulting in high rates of
water leakage.
Soon after embankment construction was completed a longitudinal crack (300 m in
length and 25 to 40 mm in width) developed on the crest. Several deep test pits were
excavated within the core to investigate the cracking, and Sherard (1973) describes the
findings. An important observation was that the no cracks were visibly evident in the
pit excavated within the upstream portion of the core, yet numerous horizontal cracks
were exposed in the pit excavated in the downstream portion of the core. The greater
number of horizontal cracks observed in the downstream portion of the core is
considered to indicate that differential settlements at the downstream interface of the
core were large. This suggests that, whilst lateral spreading is contributing to the
greater settlement of the core, the deformation of the core is also controlled to some
Chapter 7: Deformation behaviour of embankment dams Page 7.214

extent by the settlement of the upstream shoulder. The upstream orientation of the core
would also contribute to the large differential settlement with the downstream shoulder.

Figure 7.122: Djatiluhur dam; (a) location of crack observed during construction,
January 1965; and (b) deformation of monuments at elevation 103 m, January to April
1965 (Sherard 1973).

7.10.2.2 Collapse Compression on First Filling and its Influence on the Deformation
Behaviour of Cores of High Undrained Strength and Low Compressibility.
The deformation behaviour during first filling where the settlement of the upstream
shoulder (due largely to collapse compression) is much greater than that of the core is
observed for a number of case studies within the database. The core types for these
embankments include:
• Dry placed and well-compacted clay cores of medium to thick width. Examples
include Eppalock and Eildon dams. At Eppalock dam very large settlements of the
upstream slope occurred on first filling compared to much smaller settlements of the
Chapter 7: Deformation behaviour of embankment dams Page 7.215

crest (Figure 7.123), close to five times smaller when compared on a percentage
height basis.
• Embankments with well-compacted silty sand to silty gravel cores. Examples
include Cherry Valley, Cougar, Round Butte, Mud Mountain and LG-2 dams, which
range in core width from thin to thick.
• Embankments with well-compacted clayey cores of thick width placed at close to
Standard optimum moisture content. This would include the series of dams owned
by the Tennessee Valley Authority including South Holston, Watuaga and Nottely
dams (Leonard and Raine 1958).

For the silty sand to silty gravel cores, the magnitude of crest settlements (as a
percentage of dam height) on first filling generally decreased with increasing core
width. For the clayey cores, the magnitude of crest settlement was greater for the core
placed close to Standard optimum than for cores placed dry of Standard optimum,
possibly reflecting the likely lower undrained strength of the wetter placed cores and a
larger component of plastic type deformation.

0 195

190
100

Reservoir level (m. AHD) .


185
200
Settlement (mm) .

180
300

175

400
170

500
165

600
160

700 155
0 1 2 3 4 5
Time (years since end of construction)
SMP SS2 - 22m upstream SMP CS2 - crest SMP SS8 - 18m downstream
SMP SS11 - 35m downstream Reservoir

Figure 7.123: Eppalock dam, post construction settlement of SMPs on the crest and
slopes for the first five years after construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.216

The differential settlement between the upstream shoulder, due to collapse compression
on wetting, and the core usually results in the observation of longitudinal cracking on
the crest and visibly greater settlement of the upstream edge of the crest. But, this is
dependent to some extent on the embankment zoning geometry. In some embankments
the surface expression of the differential settlement may be masked within the riprap
zone on the upper upstream slope. Examples of the surface expression of differential
settlement on first filling are:
• At the 100 m high Cherry Valley dam (Lloyd et al 1958) longitudinal cracking with
differential settlement across the crack was observed along the crest at the junction
between the core and transition at both the upstream and downstream edges, as
idealised in Figure 7.30. On first filling the amount of cracking was greater on the
upstream edge of the crest.
• On first filling at the 159 m high Cougar dam (Pope 1967) longitudinal cracks with
vertical offset along the crack were located at the upstream and downstream edges
of the core. For both crack locations, the vertical settlement on the upstream side of
the crack was greater, possibly indicating some localised shear development in the
narrow core (refer Section 7.10.5, item b).
• At the 160 m high La Grande No. 2 dam (Paré 1984) longitudinal cracking over a
length of 350 m was observed along the upstream edge and centre of the crest.
Visibly greater settlements up to 300 mm were evident along the upstream edge of
the crest. The differential settlement to upstream across the crack in the centre of
the core reached a maximum of 500 mm and Paré (1984) indicates the crack to be
associated with a shear surface developed within the core.
• At the 128 m high Mud Mountain dam longitudinal cracking was observed along the
crest. Cary (1958) comments that the cracking was associated with the differential
settlement between the dumped and sluiced rockfill and the well-compacted core.
• On first filling at the 134 m high Round Butte dam longitudinal crest cracking
developed in the centre of the crest above the core to a maximum length of 150 m
and width of 15 mm (Patrick 1967). There was also some indication of differential
movement between the core and upstream transition due to differential settlement
and lateral spreading.
• At the 97 m high Watuaga dam longitudinal cracking was observed on the edge of
the crest coincident with the upstream and downstream edges of the core due to the
Chapter 7: Deformation behaviour of embankment dams Page 7.217

greater settlement of the rockfill than the core, approximately 150 to 200 mm greater
(Leonard and Raine 1958).
• At the 56 m high Nottely dam Leonard and Raine (1958) describe the longitudinal
cracking on the upstream side of the crest coincident with the upstream edge of the
core as severe. The vertical difference in settlement at the crest was greater than
250 mm.

Much of the observed longitudinal crest cracking on first filling occurs along the
upstream edge of the crest, or along both edges indicating both the upstream and
downstream shoulders settle relative to the core. In several cases, particularly at LG-2,
the crack development was indicative of shear type deformation in the core. This is
discussed further in Section 7.10.3.

7.10.3 DEVELOPMENT OF SHEAR SURFACES WITHIN THE EARTHFILL CORE

Included in the database are a number of zoned earth and rockfill embankments for
which shear surfaces (at least one, possibly multiple) are known to or thought to have
developed in the earthfill core. The timing of the initial shear surface development can
be during construction (refer Section 7.9.4), on first filling (e.g. LG-2 dam) or post first
filling. Case studies where shear deformation was evident or considered to have most
likely occurred are discussed in the following sub-sections. The summary is generally
brief and further details are given in Appendix G for most case studies.
For a number of these embankments, it is “abnormal” trends in the deformation
behaviour post first filling for which the a case study has often been identified as clearly
“abnormal” or potentially “abnormal” compared to other case studies, and in a number
of cases has involved shear type deformations in the core. These trends are summarised
in the case study discussions.

7.10.3.1 Shear Surface Development in the Core on First Filling

The deformation behaviour on first filling at LG-2, Ataturk and Copeton dams is
considered to be indicative of shear type development in the core of the embankment
during first filling. At Blowering dam a shear surface is thought to have initially
Chapter 7: Deformation behaviour of embankment dams Page 7.218

developed in the core during the latter stages of construction, with further movements
along the shear surface on first filling.

(a) Blowering dam (Section 6 of Appendix G)


A detailed analysis of the deformation behaviour at Blowering dam, including finite
difference analysis of the embankment construction, is presented in Section 6 of
Appendix G. A summary of the findings is presented here.
Blowering dam (Figure 7.124) is a 112 m high central core earth and rockfill
embankment that was constructed in the mid to late 1960’s. The medium width core
was of well-compacted medium plasticity clayey sands to sandy clays placed at
moisture contents ranging from slightly dry to slightly wet of Standard optimum (see
below). Either side of the core the filter / transition zones were of well-compacted
gravels. The rockfill comprised slightly weathered to fresh quarried phyllite, meta-
siltstone and quartzite. Placement of the weaker phyllites was limited to the outer Zone
3B. The rock types were susceptible to large loss in unconfined compressive strength
when wetted (35 to 62% reduction) and as a consequence high volumes of water were
used during placement to offset as far as practicable collapse type settlements post
construction. The Zone 3A rockfill was placed in 0.9 layers and the Zone 3B in 1.8 m
layers, and compacted by 4 passes of an 8.1 tonne smooth drum vibratory roller.

Figure 7.124: Main section at Blowering dam (courtesy of NSW Department of Public
Works and Services, Dams and Civil Section).

The moisture content specification for the core was adjusted at various stages during
construction as follows:
Chapter 7: Deformation behaviour of embankment dams Page 7.219

• Initially the specification was for placement in the range 1.3% dry to 0.7% wet of
Standard optimum moisture content (OMC), and averaged 0.3% dry.
• When the embankment height was about 33 to 37 m the specification was adjusted
to 0.7% dry to 1.3% wet of OMC, and averaged 0.3% wet.
• When at about 67 to 74 m the specification was adjusted to 1.0% dry to 1.0% wet of
OMC, and averaged 0.1% wet.

During the latter stages of construction very high vertical strains were measured
within the core at 60 to 71 m depth below crest level (between cross-arms 13 and 14 in
IVM A) as shown in Figure 7.125. During construction of the last 17.5 m to crest level
the vertical strain at this depth range increased from 4.3% to 11.9%, which was far in
excess of the magnitude of strain at other cross-arm intervals in the wet placed region of
the core. In addition, the stress-strain trend for this cross-arm interval is “abnormal”
when compared to that observed in other embankments (Figure 7.20). Development of
a shear surface within the core was considered a possible explanation for the
deformation behaviour.
In the early stages of first filling shortly after the end of construction very high
vertical strains were concentrated between cross-arms 13 and 15 (Figure 7.126). A
constriction or kink in the inclinometer tube developed below cross-arm 14 shortly after
construction and several months after the start of first filling, when the reservoir level
was still 30 to 35 m below full supply level, the measuring torpedo became blocked
between cross-arms 13 and 14. The deformation behaviour was considered to indicate
further movement on the shear surface in the core. Collapse compression of the
upstream rockfill on first filling was thought to trigger the addition deformation, either
due to development of high shear stresses at the upstream interface between the filters
and rockfill as a result of differential settlement, or due to reduction in the lateral
stresses acting on the upstream face of the core.
In 1982/83 (14 to 15 years after construction) the reservoir was subjected to a large
drawdown of 57 m to an elevation more than 70 m below full supply level. This was
the largest drawdown in the dam’s history. On drawdown, acceleration in the rate of
settlement of SMPs on the upstream shoulder and crest was measured, with a crest
settlement for the period of about 60 to 80 mm. The internal vertical settlement of the
core either side of the drawdown (only the upper 17 cross-arm intervals could be
measured due to the earlier constriction) indicated that the upper 55 m of the
Chapter 7: Deformation behaviour of embankment dams Page 7.220

embankment virtually settled as a block (i.e. only 3 mm cumulative settlement in the


upper 55 m), indicating most of the settlement on drawdown occurred in the lower 50 to
55 m of the core. It is possible that most of this settlement represents shear type
deformation on the existing shear surface reactivated on large drawdown. The trigger
for the movement was possibly the reduction in lateral stress acting on the upstream
face of the core on drawdown.
Further acceleration in the rate of settlement of SMPs on the upstream slope
occurred during the next large drawdown of 53 m in 1997/98 (29 to 30 years after
construction).

0%

2%
Vertical strain (%) .

4%

6%

X-Arm 1 to 3 (h/H = 0.97)


8% X-Arm 4 to 5 (h/H = 0.90)
X-Arm 9 to 12 (h/H = 0.74)
X-Arm 13 to 14 (h/H = 0.64)
X-Arm 15 to 17 (h/H = 0.59)
10% X-Arm 22 to 24 (h/H = 0.40)
X-Arm 30 to 31 (h/H = 0.18)

12%
0 200 400 600 800 1000 1200 1400 1600 1800
Estimated total vertical stress from numerical analysis (kPa)
Figure 7.125: Blowering dam, vertical strain during construction for selected cross-arms
intervals in the core.

(b) Ataturk dam (Section 1.1 of Appendix G)


Ataturk dam, Turkey (Figure 7.127) is a 184 m high central core earth and rockfill dam
constructed in the late 1980’s. The medium width central core is of high plasticity,
reasonably to well compacted clays to sandy clays placed at moisture contents on
average 1.5% dry of Standard optimum. Rockfill was placed in 0.6 m to 1.5 m layers
and compacted by vibratory rollers (Cetin 2002), but no specific details are available for
each zone. Moisture contents at placement were in the range 2 to 6%. The inner
upstream rockfill consisted of weathered, vesicular basalt and outer upstream rockfill of
sound basalt. The downstream shoulder consisted of sound basalt and an encapsulated
zone of pliacated limestone having a sand to clay sized fraction of 50%. Poor quality
Chapter 7: Deformation behaviour of embankment dams Page 7.221

materials were therefore used for the inner rockfill zones up and downstream of the
core.

10
2 Apr 1968
20

Depth below crest (m) . 30

40

50

60

regions of localised high


70 strain

80 15 Aug 1968 - 0.36 yrs after EOC


1 Sept 1968 - 0.40 yrs after EOC
10 July 1969 - 1.3 yrs after EOC
90
1 Apr 1970 - 2 yrs EOC, 0.4 yrs EOFF
EOC = end of construction
100 EOFF = end of first filling

foundation level
110
0 200 400 600 800
Settlement (mm)
Figure 7.126: Blowering dam, post construction internal settlement of the core from
IVM A during first filling.

Figure 7.127: Cross section of Ataturk dam (Cetin et al 2000)


Chapter 7: Deformation behaviour of embankment dams Page 7.222

The post construction crest settlement at Ataturk dam was of very large magnitude.
More than 7 metres settlement was measured in less than 7 years (close to 4% of the
dam height) and clearly stands out as “abnormal” in comparison to similar type
embankments (Figure 7.46a and Figure 7.53a).
Cetin et al (2000) indicate that large settlements occurred in June to December 1990
and again in early 1992 (1.5 to 2 years post construction) during periods of relatively
rapid rise in reservoir level to new high levels. In the early stages of reservoir filling in
June 1990, several months before the end of construction, a number of internal
monitoring gauges in the lower elevations were lost. Cetin et al (2000) also refers to
“landslides” occurring in the upstream slope in May 1992. It is possible that they are
referring to the surface expression of differential settlement between the upstream
shoulder and core. During reconstruction of the upper 6 to 7 m of the crest in 1997
slickensided surfaces were observed in the core at close to the interface between the
core and downstream filters.
Cetin et al (2000) considered slaking of the vesicular basalt in the upstream shoulder
and poor placement of the core to be significant factors in the very large settlement of
the crest post construction. Degradation of the basalt has since been discounted as a
possible cause of the large deformation (Riemer 2001).
It is difficult to surmise the potential cause/s and mechanics controlling the
deformation behaviour of the embankment given the limited information available.
Notwithstanding this, it is suspected that collapse compression of the upstream rockfill
on wetting is likely to be a significant factor. On first filling it is suspected that very
large settlements occurred in the upstream shoulder due to collapse compression on
wetting, the weathered basalt rockfill possibly being particularly susceptible. The
observations of differential settlement between the upstream shoulder and core, and the
very large magnitude of settlement of the crest tend to confirm this.
The most likely explanation for the very large settlement of the crest, large
differential settlement between the crest and downstream shoulder (Figure 7.128), and
observation of slickensided surfaces in the core at close to its downstream interface is
considered to be the development of a shear surface and shear deformations in the core
toward upstream. The available information would suggest that a shear surface (or
surfaces) formed within the core during a rising reservoir condition in the early stages of
first filling and prior to the end of construction, and that further shear type deformations
occurred during the early part of 1992 on a rising reservoir. It is possible that localised
Chapter 7: Deformation behaviour of embankment dams Page 7.223

instability developed in the core due to high shear stresses at the upstream interface as a
result of the collapse compression of the upstream rockfill on wetting.
How extensive the surface of rupture might be is not known, but the loss of internal
instruments in the lower elevations of the embankment may be related to shear
displacements, indicating the shear surface (or surfaces) is at depth.

1000

2000
Settlement (mm) .

3000

CSP 960 - Crest


4000 SMP 133 - 54m downstream
SMP 134 - 115m downstream
5000

6000

7000

8000
0 1 2 3 4 5 6 7 8
Time (years since end of construction)
Figure 7.128: Post construction settlement of the crest and downstream shoulder at
Ataturk dam.

(c) La Grande No. 2 dam (Section 1.12 of Appendix G)


The LG-2 dam in Quebec is a 160 m high central core earth and rockfill dam, with the
core slightly inclined to upstream, that was completed in October 1978. The thin core is
of well-compacted non-plastic gravelly silty sand moraine deposits and supported by
moderately wide and well-compacted filter / transition zones of gravelly sands to sandy
gravels. The rockfill shoulders are of quarried granitic gneiss dry placed in 0.9 to 1.8 m
thick layers and compacted with 4 passes of a 9 tonne smooth drum vibratory roller.
The reservoir was filled to close to maximum water level over the period October
1978 to December 1979 (0.1 to 1.2 years after construction). Since first filling the
reservoir operation is not known.
Extensive cracking of the crest, described by Paré (1984), occurred during the latter
stages of first filling. One of the longitudinal cracks was located close to the centreline
of the crest, across which differential settlement to upstream was about 500 mm after
first filling (Figure 7.129). Investigation undertaken after first filling found that this
Chapter 7: Deformation behaviour of embankment dams Page 7.224

crack was near vertical and in the order of 150 to 200 mm wide decreasing to 50 mm
wide at 3.5 m depth. Paré (1984) also refers to a “sharp tilt” that developed in
September 1980 within an inclinometer located in the upstream portion of the core
(Figure 7.129), which became blocked at about 18 m depth in November 1980. The
timing of these observations is 1.9 to 2.2 years after end of construction, almost 1 year
after completion of first filling.

Figure 7.129: LG-2 dam; crack in crest and possible shear plane in core (Paré 1984)

Paré (1984) considered the localised straining in the inclinometer as development of a


shear plane in the core. He attributes the longitudinal cracking and shear formation to a
combination of the large downstream displacement of the core on first filling and
collapse settlement of the upstream rockfill on wetting.
An interesting aspect of the deformation behaviour is the timing of the “sharp tilt”
and blockage in the inclinometer. Paré (1984) indicates that this occurred almost 1 year
after completion of first filling. But, the shear surface in the core developed during the
latter stages of first filling as indicated by the timing of the crack and the large
settlement (close to 600 mm) of the upstream edge of the crest on first filling. It would
have been expected that some indication of the shear formation would have been
identified in the deformation of the inclinometer during first filling, however, there is no
indication from Paré (1984) that any tilt was recorded during first filling. Maybe the
inclinometer was not installed until after first filling when a potential slip surface was
identified.
The likely mechanism of the initial shear formation in the core is considered to be a
result of high shear stresses on the upstream interface of the core that developed due to
the differential settlement between the core and the upstream rockfill shoulder. The
Chapter 7: Deformation behaviour of embankment dams Page 7.225

observation of further shear deformation from September to November 1980 is possibly


drawdown related and due to the reduction in lateral support on the upstream face of the
core as the reservoir level was lowered. This reduction in lateral support possibly led to
a locally unstable condition of the already sheared upstream wedge of core, which then
deformed to upstream until adequate lateral support was provided by the upstream
shoulder.

(d) Copeton dam (Section 1.5 of Appendix G)


The 113 m high Copeton dam in New South Wales, Australia (Figure 7.130) is a central
core earth and rockfill dam that was constructed in the early 1970’s. The medium width
core is of well-compacted clayey sands of medium plasticity placed at moisture contents
in the specified range of 1% dry to 1% wet of Standard optimum. The low pore water
pressures that were developed during construction suggest the core was placed on the
dry side of Standard optimum. The rockfill shoulders are of quarried granite placed in
1.2 m (Zone 3B) to 3.7 m (Zone 3C) thick layers and compacted with 4 passes of a 9
tonne smooth drum vibratory roller. No water was added during construction.

Figure 7.130: Main section at Copeton dam (courtesy of New South Wales Department
of Land and Water Conservation)

The post construction deformation behaviour of the SMPs on the crest and slopes of the
embankment was “normal” in comparison to similar type embankments. However, the
internal core settlement within the internal settlement gauges located slightly upstream
of dam axis (IVM A and IVM C) indicated the development of a possible shear zone in
Chapter 7: Deformation behaviour of embankment dams Page 7.226

the core at close to its upstream interface. The internal deformation of the core during
construction in these IVMs was considered “normal” indicating the shear surface did
not develop until first filling.
Figure 7.131 shows the development of localised zones of high strain in IVM A,
located upstream of the dam axis, at depths of 20 to 30 m below crest level. Similar
localised zones of high strain were measured in IVM C at 20 m depth below crest level.
In both IVMs the highest region of strain is between the two top cross-arms and close to
the upstream interface of the core with the upstream filter as shown in Figure 7.130.
The regions of high shear strain are at elevations where the Zone 3C rockfill (dry placed
in 3.7 m lifts) is located immediately upstream of the Zone 2 filter. By March 1999 post
construction vertical strains between the upper cross-arms was 6.3% in IVM A (cross-
arms 59 to 60) and 8.5% in IVM C (cross-arms 43 to 44), or 95 and 129 mm
respectively. In contrast the settlement profile at IVM B, located downstream of the
dam axis, shows no localised region of high strain.
The settlement versus time plot of the upper cross-arms intervals in IVMs A and C
(Figure 7.132) shows that increases in settlement between the cross-arms occur at
similar time periods; during first filling, sometime between 9 and 13 years, and
sometime between 20 and 26 years. The localised settlement in the latter periods
possibly occurs during rising reservoir level. This is confirmed from the settlement
records of SMPs on the upstream edge of the crest which show a small but perceptible
increase in the rate of settlement between 10.35 to 11.2 years post construction
coincident with the rise in reservoir level to full supply level.
The possible cause and mechanism associated with the development of the localised
regions of high strain can only be surmised from the data. It is reasonable to conclude
that localised straining between the upper cross-arms in IVMs A and C is concentrated
during periods of rising reservoir level, most likely when raised above the elevation of
these regions of high strain at 545 to 560 m. A likely explanation is that a shear zone
developed within the upstream region of the core on first filling due to high stresses at
the upstream core / filter / rockfill interface developed from differential settlement
associated with collapse compression on wetting of the Zone 3C dry placed and poorly
compacted rockfill. It is notable that the regions of localised high strain in the core are
located 5 to 15 m above the elevation where the Zone 3C rockfill was placed
immediately upstream of the upstream filter.
Chapter 7: Deformation behaviour of embankment dams Page 7.227

The subsequent shear displacements during rising reservoir at about 10.5 to 11


years, and then again at 22 or 25 years, are most likely due to further differential
settlement at the upstream interface as indicated by the SMP and IVM data records. A
possible explanation for the differential settlement post first filling is softening or
degradation of the upstream rockfill over time.
The presence of transverse and longitudinal cracking in the bitumen seal on the crest
(first observed in June 1977, 4 years after construction), and the “visually evident”
greater settlement of the upstream side of crest noted in surveillance reports (Land and
Water Conservation NSW 1995a) give further support to the mechanism of shear
development.

0
27 June 1973
(0 yrs post EOC) localised zones of
10 high strain

20

30
Depth below crest (m) .

40

50

60

70

80

Apr 1974 - 0.85 years post EOC


90 Nov 1978 - 5.4 years post EOC
Aug 1986 - 13.1 years post EOC
Mar 1999 - 25.7 years post EOC
100 EOC = end of embankment construction
foundation level
110
0 100 200 300 400 500 600 700 800
Settlement (mm)

Figure 7.131: Copeton dam, post construction internal settlement profile in the core at
IVM A.
Chapter 7: Deformation behaviour of embankment dams Page 7.228

140 574

564
120
Settlement between X-arms (mm) .

554

Reservoir Level (m, AHD) .


100 544

534
80

524
60
514

40 504
IVM C, X-arm 43 to 44 (21.0 m below crest, h/H = 0.25)
IVM A, X-arm 59 to 60 (21.4 m below crest, h/H = 0.20) 494
20 IVM A, X-arm 53 to 55 (29.7 m below crest, h/H = 0.28)
484
Reservoir

0 474
0 5 10 15 20 25 30
Time (years since end of construction)

Figure 7.132: Copeton dam, post construction settlement between cross-arms versus
time.

7.10.3.2 Shear Surface Development in the Core Post Filling

There are a number of case studies for which a shear surface has or is thought to have
developed in the core post first filling. At Eppalock and El Infiernillo dams internal
monitoring records indicate that a shear developed in the core, at Djatiluhur dam it is
most likely and for Eildon, Bellfield, Cougar and several other dams possibly
developed. Details for each of these case studies are presented in Appendix G
referencing the sources of data, and they are summarised below (except for Cougar dam
which is summarised in Section 7.10.5).
As previously indicated, for virtually all but Eppalock and Djatiluhur dams the
overall stability of the embankment is not in question. Therefore these two cases will be
dealt with first and in more detail.

(a) Eppalock Dam (Section 1.10 of Appendix G)


The 47 m high Eppalock dam (Figure 7.123), located in central Victoria, Australia is a
central core earth and rockfill embankment that was constructed in the early 1960’s.
The central core of medium plasticity sandy clays were placed in 380 mm loose
thickness layers and compacted by sheepsfoot rollers at moisture contents on average
0.8% dry of Standard optimum. The Zone 2A gravel filters were lightly roller and the
Chapter 7: Deformation behaviour of embankment dams Page 7.229

crushed basalt rock Zone 2B filter zone was end dumped in high lifts. Rockfill was of
quarried basalt was dry placed in 2 to 4 m lifts (and one 10 m lift) and spread by tractor.
The reservoir was first filled over the period from May 1962 to November 1963 (0.2
to 1.65 years after construction) and since then is subjected to a seasonal drawdown of
typically 3 to 5 m. Larger drawdowns of 7 to 10 m occurred at 5 to 6 years (1967/68),
14 to 16 years (1976/78), 20 to 21 years (1982/83), 32 to 33 years (1994/95) and 36 to
38 years (1997/99) after construction.
On first filling large settlement of the upstream shoulder occurred due to collapse
compression on wetting of the dry placed and poorly compacted rockfill. Similar large
settlements also occurred for the downstream shoulder in the first 3 to 4 years after
construction. Comparatively, the magnitude of crest settlement was much smaller
(Figure 7.134). Internal settlements within the core (measured in the internal settlement
gauge) over this period were “normal”.
During the larger drawdowns several “abnormal” trends were evident in the
deformation behaviour, including:
• Accelerations in the settlement rate of SMPs on the crest, in particular during the
drawdowns at 20 to 21 years (SMP CS1 only), 32 to 33 years and 35 to 38 years
after end of construction (Figure 7.133). It is notable that the influence of the first
large drawdown in 1967/68 (5 to 6 years) on the crest settlement is negligible, but
then much larger settlements occurred during the drawdown at 20 to 21 years and
increasing settlement magnitude during later large drawdowns. This deformation
pattern suggests softening of material strength parameters with time.
• The non-recoverable upstream crest displacement on large drawdown at 32 to 33
years and 37 years (Figure 7.135).
• Acceleration of settlement and non-recoverable upstream displacement of SMP SS2
on the upstream shoulder.
Chapter 7: Deformation behaviour of embankment dams Page 7.230

Figure 7.133: Eppalock dam, original design at maximum section (Woodward Clyde 1999)
Chapter 7: Deformation behaviour of embankment dams Page 7.231

195
a) 0

100 190

Reservoir level (m. AHD) .


200
SMP CS1 185
Settlement (mm) .

300
180

400
175
500

170
600

165
700

800 160

900 155
0 5 10 15 20 25 30 35 40
Time (years since end of construction)
SMP SS2 - 22m upstream SMP CS2 - crest SMP SS8 - 18m downstream
SMP SS11 - 35m downstream SMP CS1 - crest Reservoir

0.0%
b)
drawdown at 20 to 21 yrs

d'down at 32/33 yrs


Settlement (% of height from .

0.5%
d'down at 37 yrs
SMP to foundation) .

1.0%

end of
1.5%
first filling

2.0% SMP SS2 - 22m upstream


SMP CS2 - crest
SMP SS8 - 18m downstream
SMP SS11 - 35m downstream
SMP CS1 - crest
2.5%
0.1 1.0 10.0 100.0
Time (years since end of construction)

Figure 7.134: Eppalock dam, post construction settlement at main section and CS1.

Inclinometers were installed in the crest of the embankment in 1997. On drawdown in


1998 (35 to 36 years) a localised shear type displacement of 1 to 2 mm was observed at
11 m depth below crest level in the inclinometer located next to SMP CS1. The
localised displacement occurred between mid March and mid April 1998 when the
reservoir was drawn down below 186.8 m AHD to a low at 186.4 m AHD. Further
localised shear type displacement at this depth and also at 4 m depth were detected
Chapter 7: Deformation behaviour of embankment dams Page 7.232

during (and following) placement of additional rockfill on the upper berm of the
upstream slope at the time of the remedial works in 1999. Davidson et al (2001) refer to
observation of shear type deformations along existing longitudinal cracks in the core
and to the surface expression of the shear type deformation on the downstream batter of
the exposed core during the remedial works.
These observations are clear evidence of the presence of a surface of rupture within
the core oriented to upstream. The surface of rupture has formed because of a lack of
support of the relatively stiff core by the rockfill and is not indicative of an overall low
factor of safety, which was about 1.4. The accelerations in deformation of SMPs on the
crest and upstream slope during large drawdown at 20 years (1982/83) are possibly the
first indication of shear development in the core. Significant shear type deformations
are unlikely to have occurred before this because they are not evident in the larger
drawdown at 5 to 6 years (1967/68) and nor are they evident in the internal deformation
of the core (IVM 1) during construction and the first three years after construction. The
increasing magnitude of crest settlement and lateral spread of the shear zone in the core
on subsequent large drawdown at 33 years and 37 years is considered to be indicative of
a gradual softening and possible reduction in factor of safety with time. In comparison
to other embankments this trend of increasing magnitude of deformation on subsequent
similar sized drawdowns is unusual.
It is notable that no indication of shear development was observed in the first 3 to 4
years after construction when the settlement of the rockfill shoulders was very much
greater than the core. After this period the settlement of the core has been of similar
magnitude to that of the rockfill shoulders. The longitudinal cracking on the crest
observed from 1973 (11 years after construction) is therefore unlikely to be caused by
the differential settlement between the shoulders and core alone, although it is a
contributing factor that led to the initial crack development, softening of the core and
subsequent shear type deformations within the core. The initial cause of the cracking is
not precisely known, but the ongoing cracking is considered to be reflective of
differential deformation between the upstream and downstream portions of the core, and
not just as shear type deformations. There are not the monitoring records to confirm
such behaviour, but the change in displacement trend of SMP CS2, which started in
about November 1975 (13.65 years), from being similar to that of the downstream slope
to being similar to that of the upstream slope provides some indication of the differential
deformation behaviour of the core. The timing of the change is shortly after the initial
Chapter 7: Deformation behaviour of embankment dams Page 7.233

observation of longitudinal cracking in 1973 suggesting that development of significant


cracking in the core initially occurred from about 1973 to late 1975, and preceded the
shear deformation in the early 1980’s.
Another significant observation at Eppalock dam is the softening that has developed
in the core, particularly the upper 5 to 6 m, confirmed from piezocone testing and
pressuremeter testing in boreholes. Test pits and boreholes have revealed a series of
softened zones at angles of 45 degrees to near vertical extending beyond 4 m depth to
about the full supply level.

900
SMP SS2 - 22m upstream
SMP CS2 - crest
750
SMP SS8 - 18m downstream
SMP SS11 - 35m downstream
(downstream is positive) .

600
Displacement (mm) .

end of
first filling
450
d'down at 37 yrs
300

150
drawdown at 20 to 21 yrs

0 d'down at 32/33 yrs

-150

-300
0.1 1.0 10.0 100.0
Time (years since end of construction)
Figure 7.135: Eppalock dam, post construction horizontal displacement at the main
section.

(b) Djatiluhur dam (Section 1.7 of Appendix G).


The materials and placement methods of the 105 m high Djatiluhur dam (Figure 7.121)
and the surface and internal cracking in the core that developed from the latter stages of
construction were previously discussed in Section 7.10.2.1, item e.
The post construction settlement of the crest (Figure 7.136) shows acceleration in
the settlement rate of the core and upstream edge of the crest occur post first filling on
large drawdown to below about elevation 80 m. In comparison to other embankments
that show acceleration in the settlement rate on large drawdown, the magnitude of
settlement during drawdown at Djatiluhur dam is large (in the order of 120 to more than
300 mm). In addition, the magnitude of settlement of the core between the drawdowns
in 1972 (7 years) and 1982 (17 years) is similar, which is an unusual observation and
Chapter 7: Deformation behaviour of embankment dams Page 7.234

has only been observed at Djatiluhur and Eppalock dams. The general trend is for either
negligible or reduced magnitude settlement on a second large drawdown of similar
magnitude, or a larger magnitude second drawdown is required for an increase in the
settlement rate.
Sowers et al (1993) comment that the continuing deformation of the central and
upstream region of the crest is reflective of the “highly” stressed state of the upstream
slope as indicated by its marginal factor of safety under static loading. They add that
the accelerations in settlement on large drawdown are potentially indicative of shear
type displacements.
The writer agrees with Sowers et al (1993). The settlement data on drawdown also
indicates a softening in the material strength properties over time. This would suggest
that the development of a shear surface and strength loss due to shearing on this surface
of rupture, at least in the core.

3000 110

100

2500
90

Reservoir Level (elevation, m) .


Settlement (mm) .

80
2000

70

1500 60

50

1000
40

centre of crest (main section) 30


500 SMP C7b - upstream edge of crest
SMP C9 - upstream edge of crest
20
Reservoir

0 10
0 5 10 15 20 25 30 35
Time (years since end of construction)

Figure 7.136: Djatiluhur dam, post construction crest settlement.

(c) Bellfield Dam, Victoria, Australia (Section 1.2 of Appendix G).


Bellfield dam is a 55 m high central core earth and rockfill dam constructed in the mid
1960’s. The medium width core of sandy clays to clayey sands was compacted in 380
mm layers (loose thickness) at a specified moisture content range from 1.5% dry to
1.5% wet of standard optimum. The thin filter zones were compacted with steel flat
Chapter 7: Deformation behaviour of embankment dams Page 7.235

drum rollers and the rockfill shoulders dry placed/dumped in 1.2 to 9.1 m lifts sloped at
the angle of repose. The rockfill was sourced from mostly quarried sandstones but with
some siltstones and mudstones.
The design and construction methods are very similar to those at Eppalock dam.
The data records obtained (Snowy Mountains Engineering Corp. 1998a) for the
SMPs and IVM in the core only cover the post construction period from 1987 to 1997
(21 to 31 years after construction). The IVM records from end of construction to
February 1987 (Figure 7.45a) show two regions of high localised vertical strain within
the core, one at 14 to 15 m below crest level and the other at 35 m depth. The timing
and cause of the concentrated settlement is not known. It may represent local yielding
or possible localised shear type displacements in the core.
The records from February 1987 to November 1997 (Figure 7.45b) show a localised
zone of higher vertical strain developing at 28 to 30 m below crest level, inconsistent in
elevation with those developed prior to 1987.

Figure 7.137: Bellfield dam, main section at chainage 701 m (courtesy of Wimmera
Mallee Water)

(d) El Infiernillo Dam, Mexico (Section 1.9 of Appendix G).


The materials and placement methods of the 148 m high El Infiernillo dam (Figure
7.111) and the internal vertical deformation behaviour during first filling were
previously discussed in Section 7.10.2.1, item a. The post first filling internal
settlement records in the core over the period mid 1966 to 1972 (Marsal and Ramirez de
Arellano 1972), 2.6 to 8 years after construction, indicate a region of high vertical strain
developed in the core at about 45 m depth below crest level (Figure 7.138). The region
of high strain is evident in October 1967 and continued to progressively develop into the
early 1970s. Over the corresponding period the displacement shows a block type
Chapter 7: Deformation behaviour of embankment dams Page 7.236

deformation to downstream above elevation 135 m, and possibly a small reverse


displacement to upstream between elevations 130 and 135 m.
At the time of the formation of the localised region of high strain a sustained period
of accelerated settlement of the crest and downstream shoulder, and downstream
displacement was occurring (Figure 7.139). Marsal and Ramirez de Arellano (1972)
comment that the increased rate of deformation was coincident with the flooding of the
lower portion of the downstream rockfill due to high tail water levels in October 1966
(2.8 years), January 1967 (3.1 years) and September 1967 (3.8 years) and periods of
heavy rainfall. Collapse compression in the downstream rockfill due to wetting from
inundation and rainfall is suspected as the cause of the increased rate of settlement and
downstream displacement from late 1966 to 1968 of the crest and downstream shoulder.
Interpretation of the deformation behaviour and mechanism leading to development
of the localised region of high vertical strain is not clear. It may be related to the
deformation of the downstream shoulder, but it could be related to localised instability
of the core under the larger drawdowns from 1968 onward. The internal core
displacement to upstream at the location of high strain may indicate it is drawdown
related.

Figure 7.138: El Infiernillo dam, internal (a) settlement and (b) displacement in the core
from 1966 to 1972 (Marsal and Ramirez de Arellano 1972)
Chapter 7: Deformation behaviour of embankment dams Page 7.237

Marsal and Ramirez de Arellano (1972) comment that the “abnormalities” in the post
construction deformation behaviour are largely controlled by collapse type deformations
of the dry and poorly placed rockfill, and due to incompatibility of the stress-strain
characteristics between the materials used in the embankment. They add that the
deformation of the core is very sensitive to interactions with the surrounding granular
mass.

a) 1400 171

161
1200

Reservoir Level (m) .


151
Settlement (mm) .

1000 141

131
800
Mag. 7.6 earthquake
121
in March 1979
600
111

400 101
SMP M10 - crest, downstream edge
SMP M23 - 110m downstream
91
IVM I1 - Crest, centre
200
IVM D1 - 31m upstream
81
Reservoir

0 71
0 2 4 6 8 10 12 14 16 18
Time (years since end of construction)

500 171
b)
161
400

Reservoir Level (m) .


151
(downstream is positive) .
Displacement (mm) .

300 141

131
200
Mag. 7.6 earthquake
121
in March 1979
100
111

0 101
SMP M10 - crest, downstream edge
SMP M23 - 110m downstream 91
-100 IVM I1 - Crest, centre
81
Reservoir

-200 71
0 2 4 6 8 10 12 14 16 18
Time (years since end of construction)

Figure 7.139: El Infiernillo dam, post construction (a) settlement and (b) displacement
of surface markers.
Chapter 7: Deformation behaviour of embankment dams Page 7.238

(e) Eildon Dam, Victoria, Australia (Section 1.8 of Appendix G)


Eildon dam (Figure 7.140) is a central core earth and rockfill embankment of 80 m
maximum height that was constructed in the early to mid 1950’s. The central core
consisted of an inner zone (Zone 1A) of medium plasticity silty to sandy and gravelly
clays and outer zone (Zone 1B) of clayey sands to silty sands. Both zones were placed
on the dry side of Standard optimum and well compacted. The sandy gravel filter /
transition zone was generally placed by end dumping without compaction, and the
rockfill shoulders were placed in at least 2 m thick layers, probably without the addition
of water and not formally compacted. The rockfill was sourced from quartzitic
sandstone for Zones 3A and 3B, and the random rockfill zone (Zone 3C) consisted of
unsuitable rock that was poorly graded and contained a high fraction of finer sized
rockfill.
Aspects of the post construction deforma tion indicate the possibility that a shear
formed within the core sometime after the end of construction and that deformations on
the shear surface occur during periods of large drawdown. They are:
• The settlement records from the internal vertical measurement gauges (IVM)
installed in the core (Figure 7.142). The records for the period from 26 years after
end of construction show localised zones of high vertical strain (IVM ES2 and ES3)
at 17 to 20 m depth below crest level, a large portion of which occurred during the
latter stages of the large drawdown in 1982/83 at 27 years. These zones are located
in Zone 1A within 1 to 2 m of the upstream interface with Zone 1B. At IVM ES3 a
blockage or constriction in the tube has limited measurements to the upper 20 m of
the gauge.
• SMPs on the crest and upstream slope show an increase in settlement rate on large
drawdown at 13 years (1968) and 27 years (1982/83) (Figure 7.141), and also show
a non-recoverable upstream displacement at 27 years. Most of the other SMPs on
the crest and upstream slope (Snowy Mountains Engineering Corp. 1999a)
displayed a similar increase in settlement rate and non-recovered upstream
displacement during these drawdowns.

The cause of the possible shear within the core at Eildon dam is not known. It could
be due to poor support from the rockfill shoulders as in the case of Eppalock dam or due
to differential settlement between the upstream Zone 1B outer sandy loam core and the
inner Zone 1A clay core.
Chapter 7: Deformation behaviour of embankment dams Page 7.239

Figure 7.140: Main section at Eildon dam (courtesy of Goulburn Murray Water).
Chapter 7: Deformation behaviour of embankment dams Page 7.240

1600 290

1400 280

Reservoir Level (m, AHD) .


1200 270
Settlement (mm) .

1000 260

800 250

600 240

400 230

SMP SS10 - 32m upstream SMP CS7 - crest, centre


200 220
SMP SS7 - 28m downstream SMP SS4 - 54m downstream
Reservoir
0 210
0 5 10 15 20 25 30 35 40 45

Time (years since end of construction)

Figure 7.141: Eildon dam, post construction settlement of SMPs at chainage 685 m.

10

20
Depth below crest (m)

localised zones of high


30 vertical strain

Base of ES 1
40

50
IVM ES1 - chainage 73 m
IVM ES2 - chainage 226 m
60 IVM ES3 - chainage 683 m

70
Base of ES 2
H = 73.6 m Base of ES 3
H = 76m
80
0 20 40 60 80 100 120 140

Settlement (mm)
(from 1981 to 1998, 26 to 43 years post EOC)

Figure 7.142: Eildon dam, internal settlement profiles in core from IVM records for the
period 1981 to 1998.
Chapter 7: Deformation behaviour of embankment dams Page 7.241

7.10.4 POST FIRST FILLING ACCELERATION IN DEFORMATION THAT IS NOT KNOWN TO


BE SHEAR RELATED

A number of central core earth and rockfill dams show periods of acceleration in the
deformation rate post first filling that may not be related to shear deformation in the
core. Some of these have already been discussed, including the influence of earthquake
such as at Matahina dam (Figure 7.64) and El Infiernillo dam (Figure 7.139a), and the
influence of tail water impoundment of the lower downstream shoulder or heavy
rainfall, such as at El Infiernillo dam (Figure 7.139a).
In a number of other dams a short period of acceleration and small settlement of
SMPs on the crest or upstream shoulder is observed during the first and occasionally the
second drawdown. Case studies include, but are not limited to Beliche, Dartmouth,
Geehi, Parangana, Cherry Valley, Cougar, Wyangala and Gepatsch dams. The case
studies include dams where collapse compression of the upstream rockfill on wetting is
potentially significant, and others where it is less significant such as for well and
reasonably to well compacted dry placed rockfills (e.g. Dartmouth and Wyangala
dams). The deformation behaviour is not considered related to shear type deformation
within the core since there are no localised zones of high strain in the IVM data.
Although not well understood, several possible explanations for the deformation
behaviour have been considered most of which relate to softening of the strength and/or
compressibility properties of the materials on wetting.
For rockfills susceptible to collapse compression on wetting, the greatest effective
vertical stress acting on the saturated rockfill occurs immediately on wetting. Once the
reservoir level is raised slightly the vertical and lateral effective stresses in the upstream
rockfill will decrease. If the reservoir is then lowered, the maximum vertical effective
stress will increase, but will not exceed the previous maximum level at initial saturation
assuming the rockfill is free draining. With respect to the deformation behaviour after
saturation and collapse compression, the expectation is that a slight heave may occur on
decreasing effective vertical stress and a slight settlement may occur on increasing
effective stress as the stress path oscillates along an unloading / reloading path for
which the rockfill modulus would be high. A large increase in settlement on increasing
vertical stress would therefore be unexpected.
At Gepatsch dam, Schober (1967) attributed the acceleration in settlement of the
SMP on the upstream shoulder on drawdown to increased effective vertical stresses in
the upstream shoulder due to completion of the embankment construction at high
Chapter 7: Deformation behaviour of embankment dams Page 7.242

reservoir levels (refer Section 1.11 of Appendix G). This is a reasonable argument for
this case study. At Dartmouth dam this reasoning would partly explain the settlement of
the upstream shoulder on drawdown, but part of the settlement occurred before the
reservoir had been drawn down to below the reservoir level at end of construction. At
Geehi dam, where first filling did not start until after the end of construction, the
acceleration in settlement of the upstream shoulder on large drawdown cannot be
explained by this reasoning. Nor can it explain the acceleration in deformation at
Copeton and Wyangala dams.
Settlements of the crest of these dams are probably more readily explainable by
softening of the shear strength and compressibility of the earthfill. Equilibrium pore
water pressure conditions after construction can take many years to develop in the core.
It is conceivable therefore that on drawdown effective vertical stresses for the near
saturated earthfill may exceed those previously experienced on first filling and
acceleration in settlement may occur under the “softened” compressibility properties of
the earthfill. This may explain the observed increase in settlement rate on first
drawdown for the more permeable silty earthfills such as at Cherry Valley and
Parangana dams.
For the wet placed clayey earthfills (e.g. at Beliche and Dartmouth dams) softening
of the strength and compressibility properties of the earthfill is not a valid explanation
for the acceleration in deformation because the earthfills are already close to saturation.
In these cases it is likely that the reduction in hydrostatic pressures acting on the
upstream face of the core as the water level is drawn down influences the deformation
behaviour of the core. On drawdown, the reduction in hydrostatic stress will initially
cause the embankment crest to deflect upstream as total stresses are reduced in the core
and downstream shoulder. At some point though, lateral stresses must increase in the
upstream shoulder to equilibrate the lateral stresses in the wet placed core. For rockfills
where collapse compression and reduction of the compressibility occurred during first
filling, it is conceivable that this increase in lateral stress will be associated with a net
upstream displacement of the core / upstream shoulder interface, resulting in lateral
spreading of the core. For wet placed cores of low undrained strength, the lateral
spreading is likely to largely occur as undrained plastic deformation and will be
associated with vertical compression to maintain volumetric consistency, and hence
settlement of the crest.
Chapter 7: Deformation behaviour of embankment dams Page 7.243

7.10.5 OTHER CASE STUDIES WITH POTENTIALLY “ABNORMAL” DEFORMATION


BEHAVIOUR POST CONSTRUCTION

Further discussion on potentially “abnormal” aspects of the deformation behaviour at


Svartevann, Cougar and Wyangala dams is presented due to their informative nature.
The deformation behaviour relates to:
• The very large magnitude settlement and displacement of the crest and downstream
shoulder at Svartevann dam in the first four years after construction.
• The timing of the large deformation and longitudinal crest cracking at Cougar dam.
• The internal settlement of the earthfill core at close to its upstream interface at
Wyangala dam.

A summary of each case study is presented below. Additional details are provided
in Appendix G and in the references.

(a) Svartevann dam, Norway (Section 1.13 of Appendix G)


The 129 m high Svartevann dam (Figure 7.143) is a central core earth and rockfill dam
with a thin, slightly upstream sloping core. It was constructed in the seasonally warmer
months between 1973 and 1976 and stored water during construction. The core of silty
sand to silty gravel moraine deposits was placed at an average moisture content of 0.4%
wet of Standard optimum and well compacted by heavy vibrating rollers. The filter /
transition zones of gravels and crushed rock were sluiced and well compacted. Rockfill
of quarried granitic gneiss was dry placed and reasonably compacted (2 metre lifts
compacted with 8 passes of a 13 tonne smooth drum vibrating roller).

Figure 7.143: Main section at Svartevann dam (Kjœrnsli et al 1982).


Chapter 7: Deformation behaviour of embankment dams Page 7.244

The post construction deformations at Svartevann dam, particularly during the period of
first filling, were relatively high. The magnitude of the crest displacement on first
filling at 1100 mm was a clear outlier to other case studies (Figure 7.38 and Figure 7.72)
and the crest settlement and deformation of the downstream shoulder very high, but
possibly not “abnormally” so. Compared to the case study data for zoned earth and
rockfill dams with moraine cores (Dascal 1987) the crest deformation at Svartevann is
clearly very large.
As shown in Figure 7.144, a large portion of the surface deformation at the crest and
downstream slope occurred during the final 20 m raising of the reservoir to full supply
level, although significant settlements were also measured for the upper downstream
slope shortly after the end of construction. Finite element analysis by Dibiagio et al
(1982) was unable to accurately model the deformation behaviour within the
downstream shoulder, it significantly over-predicted the horizontal displacements and
significantly under-predicted the settlements.
The post construction internal deformation of the downstream shoulder (Kjœrnsli et
al 1982) shows that large settlements and displacements occurred in the mid to upper
region of the shoulder in the first 4 years after end of construction. Vertical strains were
estimated at 1.4 to 2.2% in the mid to upper region of the downstream rockfill for this
post construction period, with much lower vertical strains, 0.5 to 1.0%, estimated for the
lower 45 to 50 m (below elevation 820 m). The deformation time plots show that very
large settlements and downstream displacements of the mid to upper slopes occurred
during the latter stages of first filling, but not on the lower downstream slope.
The large crest deformations on first filling observed at Svartevann dam appear to
be related to the deformation of the downstream shoulder. In effect, the core deforms
with the downstream shoulder.
The inability of the finite element analysis (Dibiagio et al 1982) to model the
deformation behaviour on first filling, particularly the vertical component, suggests that
the application of the water load and the associated changes in stress conditions on its
own does not account for the actual deformation behaviour. In addition, time dependent
or creep related deformations would not account for the vertical deformation because
the magnitudes are too large.
It is considered that the large deformations of the mid to upper region of the
downstream shoulder are due to collapse compression in the dry placed and reasonably
compacted rockfill. The trigger for the collapse type settlement must be moisture
Chapter 7: Deformation behaviour of embankment dams Page 7.245

related, so possibly either heavy rainfall or snowmelt is the source of water and the
timing during the latter stages of first filling is somewhat coincidental. It is notable that
the accelerations in deformation rate of SMPs on the mid to upper downstream slope in
the first 2 years post construction occur at the same time period each year, from 0.0 to
0.2, 0.8 to 1.2 and 1.8 to 2.2 years corresponding to the period from June to September,
or summer. This is the seasonally wettest and warmest period for the western coast of
Norway. Therefore, heavy rainfall together with snowmelt may sufficiently wet the
downstream rockfill for collapse settlement to occur. Possibly 1978 was a relatively
wet summer or the winter one of high snowfall.

a) 1200 900

Reservoir Level (elevation, m) .


1000 875
Settlement (mm) .

800 850

600 825

400 800

200 775

0 750
0 1 2 3 4 5
Time (years since end of construction)
Crest - Station 153 (left abutment) 51m downstream (EL 870) 87m downstream (EL 845)
123m downstream (EL 820) Crest - Station 155 (main section) Reservoir
Chapter 7: Deformation behaviour of embankment dams Page 7.246

1250 900
b)

Reservoir Level (elevation, m) .


875
(downstream is positive) .
1000
Displacement, mm .

850
750

825

500
800

250
775

0 750
0 1 2 3 4 5
Time (years since end of construction)
Crest - Station 155 (main section) 51m downstream (EL 870) 87m downstream (EL 845)
123m downstream (EL 820) Reservoir

Figure 7.144: Svartevann dam, post construction (a) settlement and (b) displacement.

(b) Cougar Dam, Oregon, USA (Section 1.6 of Appendix G)


The 159 m high Cougar dam (Figure 7.145) is a central core earth and rockfill
embankment that was constructed in the early 1960’s (Pope 1967). The slightly
upstream sloping narrow core of silty gravels was placed at an average moisture content
of 1% wet of Standard optimum and well compacted. The filter / transition zones were
also well compacted. Rockfill of quarried basalt and andesite was used. Zone 3A
consisted of well-compacted sound rock, Zone 3B of sound rock placed in 900 mm
layers and tracked by 2 passes of a D8 bulldozer, and Zone 3C was of lesser quality
rock comprising up to 25% weathered rock placed in 600 mm layers and tracked by D8
bulldozer.
On first filling the post construction crest deformations at about the maximum
section, Figure 7.146, show that during the last 20 m to full supply level the settlement
and downstream displacement of the crest increased significantly in magnitude. This
latter part of the deformation was not uniform; settlements were greater at the upstream
edge of the crest and displacements greater at the downstream edge of the crest. On the
first and second drawdowns acceleration in the settlement rate of the SMP on the
upstream edge of the crest occurred, resulting in a marked increase in the magnitude of
settlement. At 3 years after construction the differential settlement of the crest had
increased to about 400 mm and the lateral spreading to about 250 mm.
Chapter 7: Deformation behaviour of embankment dams Page 7.247

Figure 7.145: Cross section at Cougar dam (Cooke and Strassburger 1988).

Cracking of the embankment was a consequence of the differential deformation


between the upstream and downstream edges of the crest. As described by Pope (1967),
longitudinal cracking of the crest was observed at the end of first filling and within
several days had extended over a length of almost 300 metres. The cracking was mostly
near the downstream core / transition interface. The cracking re-appeared in January
1965, shortly after the end of the first drawdown, located at the up and downstream core
/ filter interfaces and the downstream Zone 2A / Zone 2B interface (Figure 7.147).
Cracks widths were up to 150 mm and differential vertical displacement (to upstream)
across the crack occurred at the up and downstream edges of the core of 300 mm and
150 mm respectively.
Pope (1967) considered collapse type settleme nt of the upstream rockfill on wetting,
particularly within the lesser quality track rolled Zone 3C rockfill, to be a significant
factor in the observed deformation behaviour and cracking. But, what is interesting
with the deformation at Cougar dam is that a large proportion of the differential
deformation at the crest occurred during drawdown and not on first filling. This would
suggest that collapse settlement of the upstream rockfill on initial saturation, whilst
significant, is not the major cause of the differential deformation. A possible reason for
the large differential settlements post first filling is that the lesser quality Zone 3C
rockfill lost additional strength whilst saturated, and then under the increasing effective
stress conditions on drawdown further settlement of the upstream rockfill occurred. The
second drawdown was to a lower level than the first resulting in higher effective stress
conditions in the upstream rockfill than at the end of the first drawdown, which may
explain the further increase in settlement on the second drawdown.
Chapter 7: Deformation behaviour of embankment dams Page 7.248

The vertical offset at the downstream core / filter interface could indicate a potential
shear development in the silty gravel core caused largely as a result of the high shear
stresses at the upstream core / filter interface due to differential settlement between the
upstream shoulder and the core. The fact that no cracking was observed at the upstream
filter / rockfill interface is interesting because these materials probably have the greatest
difference in compressibility properties. It is possible that a shear surface has developed
in the upstream gravel filter at some point and as a result the upper part of the gravel
filter deforms with the upstream rockfill.

900 520
a)
800 500

Reservoir Level (elevation, m) .


700
480
Settlement (mm) .

600
460
500
440
400
420
300

400
200
TD-12 crest, upstream edge
100 TD-11 crest, downstream edge 380
Reservoir
0 360
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Time (years since end of construction)

600 520
b)
500
500
Reservoir Level (elevation, m) .
(downstream is positive) .

480
Displacement, mm .

400
460

300 440

420
200

TD-12 crest, upstream edge 400

100 TD-11 crest, downstream edge


Reservoir 380

0 360
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Time (years since end of construction)

Figure 7.146: Cougar dam; post construction (a) settlement and (b) displacement normal
to dam axis of the crest at the maximum section (adapted from Pope 1967).
Chapter 7: Deformation behaviour of embankment dams Page 7.249

Figure 7.147: Cougar dam; differential settlement and cracks on crest (Pope 1967)

(c) Wyangala dam, New South Wales, Australia (Section 1.16 of Appendix G).
Wyangala dam (Figure 7.149) is a central core earth and rockfill dam of 85 m maximum
height constructed in the 1960’s. In the deeper valley section the old concrete dam
forms the upstream toe of the earth and rockfill dam. The core (Zone 1) of silty to
clayey sands was placed at an average moisture content of 1% dry of Standard optimum
and well compacted. The filters were also well compacted. Quarried porphyritic gneiss
was the main source of rockfill and for the most part was placed without the addition of
water. Zone 3A was well-compacted in 0.9 m lifts, Zone 3B reasonably to well
compacted in 1.2 m lifts and Zone 3C reasonably compacted in 2.4 m lifts.
During construction, the deformation behaviour of Wyangala dam was considered
“normal” in comparison to other similar dams.
Post construction, the deformation behaviour of Wyangala dam was, for the most
part, considered “normal”. An interesting aspect of the deformation behaviour is the
very high vertical strains developed in the core at IVM B (Figure 7.148) between 23 and
32 m depth below crest level (1.85% at 21 years after end of construction). IVM B is
located slightly upstream of the dam axis (Figure 7.149) and the region of high vertical
strain is located near to the upstream core interface. The vertical strains in this region
are very much higher than the average vertical strain of 0.48% in IVM A at a similar
depth (located downstream of the dam axis) and in the mid to lower core region at IVM
B (average less than 0.15%). The localised high vertical strains developed steadily over
the first 11.5 years (1.2% vertical strain at 11.5 years), but increased in rate in the period
Chapter 7: Deformation behaviour of embankment dams Page 7.250

from 11.5 to 17 years (0.5% vertical strain over this period for a total of 1.7% to 17
years).
The magnitude of settlement between the cross-arms over the region of high strain
totals 155 mm in 17 years, which is approximately equivalent to the differential
settlement between the up and downstream edges of the crest (about 100 to 120 mm)
over this period plus the “normal” settlement over the depth range (estimated at 20 to 30
mm from IVM A). This would indicate that the region of high strain relates to
differential settlement between the core and upstream shoulder. Whilst this is a shear
type movement in itself, a single distinct shear surface does not appear to have
developed as indicated by the broad width of the region of high strain. It is possible that
a series of shears may have developed or that general softening of the earthfill may be a
factor within the zone of high strain.

a) 0

b) 20
10
22

20 24
Depth below crest (m) .
Depth below crest (m) .

26
30 average strain
28 of 1.85%

40
30

12 Jan 1972 (2.7 yrs post EOC)


32
50 2 July 1973 (4.2 yrs post EOC)
12 Jan 1972 (2.7 yrs post EOC)
30 May 1975 (6 yrs post EOC)
2 July 1973 (4.2 yrs post EOC) 34
5 Dec 1979 (11.5 yrs post EOC)
60 30 May 1975 (6 yrs post EOC)
36 12 April 1985 (17 yrs post EOC)
5 Dec 1979 (11.5 yrs post EOC) 24 Feb 1989 (21 yrs post EOC)
12 April 1985 (17 yrs post EOC) EOC = end of construction
38
70 24 Feb 1989 (21 yrs post EOC)
EOC = end of construction 40
80 0 50 100 150 200 250 300
Settlement (mm)
foundation level

90
0 100 200 300
Settlement (mm)

Figure 7.148: Wyangala dam, post construction internal settlement in IVM B located 9
m upstream of dam axis.
Chapter 7: Deformation behaviour of embankment dams Page 7.251

Figure 7.149: Section at Wyangala dam (courtesy of New South Wales Department of Land and Water Conservation).
Chapter 7: Deformation behaviour of embankment dams Page 7.252

7.11 “ABNORMAL” DEFORMATION B EHAVIOUR P OST CONSTRUCTION


OF EARTHFILL AND ZONED EMBANKMENTS WITH VERY BROAD
CORE WIDTHS
In Sections 7.5.3 and 7.6 the post construction deformation behaviour of a number of
rolled earthfill embankments and zoned embankments with very broad earthfill cores
were identified as “abnormal” or possibly “abnormal”. In the following sub-sections
the possible mechanisms explaining the deformation behaviour are discussed with
reference to the case study data. The association of a case study to a particular
mechanism has, in some cases, previously been identified by the owner, their
consultants or other authors, and in other cases has been considered appropriate based
on assessment of the deformation behaviour. For some cases the association is
speculative. Further details on the embankment and its deformation behaviour for most
of the case studies discussed in this section are presented in Appendix G.

7.11.1 COLLAPSE COMPRESSION OF THE EARTHFILL ON WETTING

At Hirakud dam the large settlements in the lower portion of the embankment that
occurred during construction were attributed to the effects of collapse compression of
the earthfill on wetting (refer Section 7.9.2). Two factors were significant in the
deformation behaviour; the deformation that occurred on initial wetting of the dry
placed clayey sand to clayey gravel earthfill under the stress conditions existing at the
time, and the large deformations as the embankment was constructed to design level
after the shutdown period when the wetting occurred. The large deformations during
the subsequent construction were indicative of the low shear strength and
compressibility properties of the “collapsed” and softened earthfill. The influence of
change in material strength and compressibility properties after wetting for
embankments that are initially saturated on first filling after construction is important
and its possible influence is discussed further for some of the case studies.
Rector Creek, Mita Hills and Roxo dams are considered examples of collapse
compression of dry placed earthfills that occurred due to wetting after construction.
Collapse compression possibly also occurred in the upstream shoulder to central region
of Dixon Canyon, Spring Canyon and Horsetooth dams. All these cases are discussed
further in Appendix G.
Chapter 7: Deformation behaviour of embankment dams Page 7.253

“Abnormal” deformation behaviour due to collapse compression on wetting is


generally identified by excessive settlements that largely occur on wetting. This is
particularly evident at Rector Creek, Mita Hills and Roxo dams. Another pattern
sometimes observed is an “abnormally” large upstream displacement of the crest on first
filling and later downstream displacement as the wetting front progressively develops
through the embankment.
The deformation behaviour at Rector Creek dam (Sherard et al 1963; Sherard 1973;
ICOLD 1974) presents probably the clearest example of collapse compression. The
embankment (Figure 7.150) is a 61 m high zoned earthfill embankment that was
completed in January 1947. The central earthfill region (Zones 1 and 2) was of silty to
clayey sands with low plasticity fines, the finer materials being used in the central core
region (Zone 1) and the coarser earthfills in the outer Zone 2 region. Compaction was
in 150 mm layers by heavy sheepsfoot roller and moisture contents at placement were 2
to 4% dry of Standard optimum.
The post construction deformation of SMPs on the crest near to the main section
(Figure 7.151) shows the large magnitude of crest settlement (almost 1.8% at 10 years
after construction) and the large upstream then downstream displacement of the crest.
In comparison to similar embankments the magnitude of settlement and long-term
settlement rate of the crest are “abnormally” high (Figure 7.46, Figure 7.60a and Figure
7.63), and the magnitude and direction of the crest displacement during and post first
filling clear outliers to the general behaviour (Figure 7.38, Figure 7.75 and Figure 7.77).
ICOLD (1974) consider that collapse settlement on wetting of the dry placed earthfill
and the gradual development of the phreatic surface within the embankment contributed
to the observed crest settlement and displacement behaviour, and to the observed
cracking. Initial wetting and collapse settlement of the upstream shoulder resulted in
the upstream displacement of the crest, and subsequent wetting and collapse settlement
in the central to downstream portion of the embankment resulted in the change in
direction of displacement to downstream some 2 to 2.3 years after construction. The
high long-term settlement rate of about 1.65% per log cycle of time maintained for at
least 10 years after construction is possibly reflective of the large reduction in material
strength and compressibility properties after wetting.
Chapter 7: Deformation behaviour of embankment dams Page 7.254

Figure 7.150: Rector Creek dam main section (Sherard 1953).

2.0% 200

1.8% Stn 16 - crest settlement 150


Settlement (% of dam height) .

1.6% Stn 15 - crest displacement 100

Note: displacement downstream is positive

Displacement (mm) .
1.4% 50

1.2% 0
End of first filling
1.0% -50

0.8% -100

0.6% -150

0.4% -200

0.2% -250

0.0% -300
0.1 1.0 10.0
Time (years since end of construction)

Figure 7.151: Rector Creek dam; post construction settlement and displacement versus
log time (adapted from Sherard et al 1963).

Collapse compression is considered a significant factor for the large crest settlements
measured at Roxo dam. The dam, constructed in the 1960’s, is part concrete and part
earthfill embankment. The earthfill embankment is of 27 to 32 m maximum height and
is constructed of medium plasticity clayey sands to sandy clays derived from weathered
schists placed at moisture contents in the range 2% dry of OMC to OMC (OMC =
Standard optimum moisture content). Large post construction crest settlements, up to
almost 2% of the embankment height at 8 years (Figure 7.152), were measured over a
large portion of the 80 to 100 metre long closure section of the earthfill embankment
located at the interface between the two structures. In comparison to similar
embankments, the magnitude and rate of the crest settlement is “abnormally” high
Chapter 7: Deformation behaviour of embankment dams Page 7.255

(Figure 7.46, Figure 7.60a and Figure 7.63). At 95 metres from the interface the post
construction crest settlement is more typical of “normal” type behaviour.

0.0%

0.2%
Settlement (% of dam height) .

0.4%

0.6%

0.8%

1.0%

1.2%

1.4%

Crest - 20m from buttress


1.6%
Crest - 55m from buttress
1.8% Crest - 95m from buttress

2.0%
1 10
Time (years since end of construction)

Figure 7.152: Roxo dam, settlement versus log time (adapted from De Melo and Direito
1982).

The cause of the large settlement is considered to be possibly due to collapse type
settlements on wetting of dry placed and potentially poorly compacted earthfill in the
closure section. Investigations and excavation (De Melo and Direito 1982) showed the
presence of wetted and softened layers close to those in the as placed condition, but it is
not clear to what extent these layers were observed.
De Melo and Direito (1982) do not consider collapse compression as a cause of the
large settlements. Instead, they consider it due to a combination of factors including
arching across changes in foundation geometry near to the abutment and the formation
of horizontal cracks due to differential settlement and stress transfer between the
earthfill and concrete structures. But these factors alone would only influence the
deformation in the vicinity of the interface, and, as the settlement records show, the
“abnormally” large settlements were measured more than 50 to 60 metres from the
interface. Hence, collapse compression on wetting is considered a potentially
significant factor.
Large reductions in the shear strength and compressibility properties of the earthfill
may be contributing to the “abnormally” high long-term rate of crest settlement.
Chapter 7: Deformation behaviour of embankment dams Page 7.256

At Mita Hills dam collapse compression on wetting of the outer upstream earthfill
zone (Zone 3) is considered a potentially significant factor in the “abnormally” large
magnitude of crest settlement and the upstream crest displacement on first filling
reported by Legge (1970). The embankment, constructed in the late 1950’s, is of 49 m
maximum height (Figure 7.153). The earthfill was placed in 150 mm layers and
compacted by heavy sheepsfoot rollers. The specified range in moisture content for the
outer earthfill zones was 2% dry of OMC to OMC (OMC = Standard optimum moisture
content) and 1% dry to 1% wet of OMC for the internal Zone 2 upstream of the
chimney filters.
As shown in Figure 7.60a and Figure 7.46 the magnitude of crest settlement at Mita
Hills dam of more than 1% is “abnormally” high, and the magnitude of upstream
displacement (Figure 7.75 and Figure 7.38) possibly “abnormal” in comparison to
similar type embankments. Most of the deformation occurred on first filling. Post first
filling the crest deformation has been “normal” in comparison with that of other
embankments. In addition, very high magnitude settlements were measured on the mid
abutment slopes (up to 2.3% of the embankment height at these locations) and the
deformation of SMPs on the downstream shoulder were very small (settlement of 0.27%
of the height from the SMP to foundation level at the main section) in comparison to
those on the crest.
The possibility that the deformation is related to marginal stability of the upstream
slope has been considered, however, an instability mechanism seems unlikely as no
acceleration in settlement rate or further upstream displacement is observed on
drawdown (refer Section 4.2 of Appendix G).

Figure 7.153: Main section at Mita Hills dam (Legge 1970)


Chapter 7: Deformation behaviour of embankment dams Page 7.257

For the Horsetooth Reservoir embankments Dixon Canyon, Spring Canyon and
Horsetooth dams, collapse compression is considered a possible cause of the large
magnitude of settlement of SMPs on the upper upstream slope and crest. The
Horsetooth Reservoir embankments were constructed in late 1940’s to maximum
heights of 48 to 74 m. A similar design was adopted for all of the main embankments
(Figure 7.154 is of Dixon Canyon dam) consisting of a very broad central earthfill zone
with thin outer zones of gravelly earthfill or rockfill. In the valley sections a cut-off was
excavated through the over-burden soils to bedrock with most of the embankment
founded on the over-burden soils. The cut-off design was similar for all embankments
except Horsetooth dam, where it was located upstream of the dam centreline.
The earthfill zone was constructed of mainly low plasticity sandy clays to clayey
sands and clayey gravels of alluvial origin. Finer materials were placed in the central
region of the core and coarser materials toward the outer slopes. The earthfill was
placed in 150 mm layers and compacted by tamping rollers. Moisture contents were
well dry of Standard Proctor optimum, on average 2.2% to 2.9% dry.
The post construction deformation behaviour of SMPs on the upper upstream slope
near to the main section for the Horsetooth Reservoir embankments is shown in Figure
7.155. Possible “abnormal” aspects of the deformation behaviour in comparison to
similar embankments are:
• The large magnitude of settlement of the upstream slope for Horsetooth, Dixon
Canyon and Spring Canyon dams where settlements were in the range 1.35 to 2.0%
(Figure 7.52 and Figure F2.28 of Appendix F).
• The high rate of long-term settlement of the upstream slope for Dixon Canyon dam
of 1.38% per log cycle of time (Figure 7.83). At Spring Canyon dam the settlement
rate was high over the period from 8 to 30 years, but has since reduced in the period
from 30 to 45 years.
• For Dixon Canyon dam, the large upstream displacement on and after first filling
followed by the change to downstream displacement after 30 years for the upstream
slope (Figure 7.87). The trend of displacement at Spring and Soldier Canyon dams
is similar, but at a much reduced magnitude and rate of displacement (rate per log
time).
• The crest settlement behaviour of Dixon Canyon and Spring Canyon dams is also
potentially “abnormal” in terms of the large magnitude (Figure 7.48 and Figure
7.61) and high long-term settlement rate (Figure 7.63).
Chapter 7: Deformation behaviour of embankment dams Page 7.258

Figure 7.154: Main section at Dixon Canyon dam (courtesy of U.S. Bureau of
Reclamation).

The influence of the foundation on the settlement and displacement of the crest at
Horsetooth was significant (refer Section 3.3 of Appendix G). It is also likely to have
influenced the deformation of the upper upstream slope, but to a lesser degree, as this
SMP overlies the downstream edge of the cut-off trench to bedrock. The downstream
displacement and large settlement on first filling of the SMP on the upstream slope at
Horsetooth dam (Figure 7.155) are likely due to the foundation influence.
At Dixon Canyon and Spring Canyon dams the foundation is considered to have a
limited influence on the post construction deformation behaviour and most of the
measured deformation occurs within the embankment. This is based on the comparison
of the trend of deformation at these embankments to that at Horsetooth dam where the
influence of the foundation was very significant during first filling, but much less so
post first filling. In addition, the SMPs on the upstream slope at Dixon Canyon and
Spring Canyon dams directly overlay the cut-off trench and therefore the foundation
influence will be very limited.
Collapse compression of the dry placed earthfill, particularly in the coarser soil
types likely to be present in the upstream portion of the main earthfill zone, is
considered to be a possible explanation for the observed potentially “abnormal”
deformation behaviour of the upstream shoulder of Horsetooth, Spring Canyon and
Dixon Canyon dams, and the crest at Spring and Dixon Canyon dams. The effect of
softening of the shear strength and reduction in compressibility on wetting is likely to
have some influence on the on-going very high settlement rates (rate per log time).
Chapter 7: Deformation behaviour of embankment dams Page 7.259

The similarity between the case studies considered to be susceptible to collapse


compression, including Hirakud dam, is that the materials were placed well dry of
Standard optimum moisture content, and were of low plasticity clayey sands to clayey
gravels to sandy clays. In all cases the earthfills were reportedly placed in thin layers
(150 mm) and well-compacted by rollers.

0.00%
a)
.

0.25%
(% height from SMP to foundation)

0.50%
Settlement .

0.75%

1.00%
end of first filling

1.25%

1.50% Horsetooth dam


Soldier Canyon dam
Spring Canyon dam
1.75%
Dixon Canyon dam

2.00%
1 10 100
Time (years since end of construction)

200
b)

100
.
(downstream is positive)
Displacement (mm)

end of first filling


0

-100

-200 Horsetooth dam


Soldier Canyon dam
Spring Canyon dam
Dixon Canyon dam
-300
1 10 100
Time (years since end of construction)
Figure 7.155: Post construction (a) settlement and (b) displacement of SMPs on the
upper upstream slope at the Horsetooth Reservoir embankments.
Chapter 7: Deformation behaviour of embankment dams Page 7.260

7.11.2 “HIGH” SHEAR STRESS OR MARGINAL STABILITY CONDITIONS WITHIN THE


EMBANKMENT

Three of the case studies within the database of embankment with very broad core zones
suffered failures in the upstream slope during large drawdown, these were Belle
Fourche, San Luis (Von Thun 1988; Stark and Duncan 1987, 1991) and Steinaker
(Cyganiewicz and Dise 1997) dams. The failures are clearly indicative of marginal
stability conditions of the embankments under drawdown. The possible mechanism/s of
the failures are discussed in Chapter 5.
The following discussion is on the monitored deformation behaviour of Belle
Fourche and San Luis dams, concentrating on “abnormal” or potentially “abnormal”
trends that may be potential indicators of an upstream failure condition on drawdown.
At both these embankments no SMPs or other monitoring instrumentation was located
within the actual failed section of the embankment. At Steinaker dam the failure was in
a section of the embankment that was not monitored and was different in design to the
main section, so is not discussed further. The mechanism and failure itself at each
embankment are not discussed in any detail here.

7.11.2.1 Belle Fourche dam (Section 4.1 of Appendix G)

Belle Fourche dam (Figure 7.156) is an earthfill embankment of 35 m maximum height


and about 1850 m crest length that was constructed over the period 1905 to 1911.
Foundations for the embankment consist of medium plasticity alluvial adobe clay
overlying a thin sand and gravel layer and in-turn shale bedrock. The earthfill materials
used in construction were the medium plasticity adobe clays placed in 150 mm layers,
“sprinkled” with water and compacted using heavy rollers with diagonal lugs.

Figure 7.156: Belle Fourche dam; main section as constructed in 1911 (courtesy of U.S.
Bureau of Reclamation).
Chapter 7: Deformation behaviour of embankment dams Page 7.261

The embankment design incorporated steep upstream and downstream slopes for the
type of construction materials used. The steep portion of the upstream slope was faced
with concrete slabs placed on a 0.6 m layer of gravel and supported at the toe (elevation
2920 feet) on a concrete footing and driven wooden piles. The closure section at Owl
Creek, the highest section of the embankment between stations 40+50 and 42+50, was
constructed last and was built very rapidly.
The reservoir was first filled over the period from March 1911 to September 1915
(4.2 years after construction) and is subjected to a seasonal drawdown, usually over the
late spring to early autumn period, generally in the range of 4 to 8 metres (Figure 7.158)
with occasional larger event of greater than 9 m.
The embankment has performed relatively poorly during the larger drawdown
events (Sherard 1953; USBR 1996). In 1928 (17 years after construction) longitudinal
cracking on the downstream edge of the crest (Sherard 1953) was observed in
November after the drawdown had been completed. A total of 5 cracks, each 5 to 50 m
in length, 12 to 50 mm in width and 1 to 3.5 m deep, were observed along the higher
embankment sections, including the closure section. In 1931 (19 years after
construction) a slide in the upstream slope occurred during drawdown on 2nd August.
The slide was approximately 110 m in width and 3 to 5 m in depth located within the
steeper upstream slope between Stations 43 and 46+50. The head of the backscarp was
about 7 m below crest level. The slide mass was excavated and the upstream slope
rebuilt to its original configuration in 1931.
Remedial works of the steep upstream slope were undertaken in 1939 (following the
failure of the upstream slope on rapid drawdown in 1931) and again in 1977 to address
concerns over embankment stability under drawdown. In 1939 a gravel stabilising berm
was added to the lower upstream slope below elevation 2950 feet, and in 1977 the upper
portion of the upstream slope was flattened to a slope of 2.33H to 1V, the crest widened
1.4 m to upstream and the crest re-surfaced (refer to Section 4.1 of Appendix G for a
cross section of the remediated embankment).
On large drawdown in 1985 (74 years after construction) longitudinal cracking was
observed on the crest between Stations 39 and 46, located 1.2 m from the upstream edge
of the crest. The crack, which was centred on the Owl Creek closure section, was about
200 m in length and up to 75 to 100 mm in width. A vertical displacement to upstream
of 50 mm was measured across the crack. Investigation found that the crack was
coincident with the upstream edge of the original structure, located directly above the
Chapter 7: Deformation behaviour of embankment dams Page 7.262

buried concrete kerb. On drawdown in 1988 and 1989 (77 and 78 years after
construction) the 1985 crack re-opened and the differential settlement to upstream
across the crack increased.
During piezometer installations in 1982 softened and very wet zones were
encountered within the earthfill and foundation (Hickox and Murray 1983).
Instrumentation for monitoring deformation behaviour at Belle Fourche dam is by
surface measurement points (SMP). At end of construction a series of SMPs were
installed along the crest in 1911 and were monitored for vertical deformation over a
period of about 17 years to 1928 (Figure 7.157), so no monitoring records are available
for the large drawdowns in 1928 and 1931. In 1985, some 74 years after construction,
SMPs were installed on the crest and slopes between Stations 26 and 46 and the
deformation records of the SMPs installed at Station 40 to 42 (within the Owl Creek
closure section) are shown in Figure 7.158.
When compared to similar embankments, a number of aspects of the post
construction deformation behaviour at Belle Fourche dam are clear outliers and
considered as indicative of “abnormal” behaviour. These include:
• The large magnitude of crest settlement in the first 17 years of operation, reaching
more than 2% of the embankment height (Figure 7.48 and Figure 7.60).
• The very high crest settlement rate (Figure 7.63a). At 10 to 12 years after
construction the rate was about 1.8% per log cycle of time, and increased to about
4.5% per log cycle of time at 75 to 85 years after construction.
• The very high rates of settlement of SMPs on both the upstream (Figure 7.83) and
downstream (Figure 7.78a) shoulders within the closure section.
• The very high rate of downstream displacement of the downstream slope (Figure
7.80) and upstream displacement of the upstream slope (Figure 7.87) within the
closure section at 75 to 85 years after construction.

Of the SMPs installed on the crest and slopes of the embankment after 1985, the
deformations between Stations 40 and 42 (within the Owl Creek closure section) were
up to 1.5 to 2 times greater than those measured elsewhere on the embankment, and
were very much larger at MP5 (located at Station 42+00) on the upstream slope.
During large drawdown at 74 years (1985) and again at 77 years (1988) acceleration in
settlement and non-recoverable upstream displacement is observed at MP5 on the
upstream slope. This is coincident with the observation of cracking and greater
Chapter 7: Deformation behaviour of embankment dams Page 7.263

settlement of the upstream edge of the crest. At 86 years further non-recoverable


upstream displacement was observed for MP5.
The embankment performa nce during large drawdown is indicative of the marginal
stability of the upstream slope on drawdown, as evidenced by the failure in the upstream
slope during the 1931 drawdown.
USBR (1996) attributes the observed cracking and monitored “abnormal”
deformation behaviour of the upstream slope to upstream crest region during the
drawdowns in 1985, 1988 and 1989 to settlement / consolidation of the original earthfill
under the added weight of the granular filling in the upstream shoulder berm placed in
1977. However, they do not discount the potential for deep-seated movements. A
notable aspect of the deformation behaviour of the upstream slope during large
drawdown is that only at MP5 (located within the Owl Creek closure section) is the
acceleration in settlement and displacement observed, no such behaviour is observed at
other SMPs on the upstream slope also located on the newly placed granular filling.
The deformation records after 75 years show relatively similar behaviour for the
crest and downstream slope, with average settlement rates of 6 to 10 mm/year and
downstream displacement rates of 6 to 13 mm/year. An increase in displacement rate of
SMPs on the crest and downstream slope is observed after 82 years, and is
approximately coincident with a period of higher average reservoir level. The increase
in settlement rate of the downstream edge of the crest and downstream slope after year
85 (1996) may reflect a change in effective stress conditions due to a rising phreatic
surface under the high average reservoir level.
Overall, the long-term deformation behaviour at Belle Fourche dam is “abnormal”
in comparison to similar embankments. The “abnormal” deformation of the crest and
slopes, particularly in the vicinity of the Owl Creek closure section (Stations 40 to 42),
possibly reflects the “highly stressed” conditions within the embankment due to the
steepness of the embankment slopes for an earthfill embankment constructed of medium
plasticity clays. An important aspect of the deformation behaviour is the much higher
crest settlement rates (per log cycle of time) from recent monitoring compared to those
at 10 to 12 years after construction. This possibly reflects softening of the undrained
strength properties of the earthfill due to wetting of the earthfill, strain weakening under
the high stress conditions imposed during large drawdown, lateral spreading in the crest
region and cracking in the upper portion of the embankment.
Chapter 7: Deformation behaviour of embankment dams Page 7.264

The USBR comment that the Owl Creek closure section, which was constructed
very rapidly, is a definite discontinuity along the dam embankment and that the
deformation behaviour within this closure section is somewhat unique to the rest of the
embankment. This is indicated by the greater magnitudes of deformation measured for
the SMPs on the crest and slopes of the embankment installed more than 75 years after
the end of construction.

2.5%

Station 31 - crest
Settlement (% of dam height) .

Station 34 - crest
2.0%
Station 40 - crest
MP10 - crest, downstream edge

end of first filling


1.5%

1.0%

0.5%

0.0%
0.1 1.0 10.0 100.0
Time (years since end of construction)

Figure 7.157: Belle Fourche dam; post construction crest settlement over the period
1911 to 1928 (to 17 years post construction).

a) 250 907

902
200
Reservoir Level (elevation, m) .
Settlement (mm) .

897

150

892

100

887

50
882

0 877
70 72 74 76 78 80 82 84 86 88 90
Time (years since end of construction)
MP10 - crest, downstream edge (Stn 40) MP5 - 16m upstream (Stn 42)
MP16 - 21m downstream (Stn 42) MP18 - crest, upstream edge (Stn 40+75)
Reservoir
Chapter 7: Deformation behaviour of embankment dams Page 7.265

250 907
b)
200
902

Reservoir Level (elevation, m) .


(downstream is positive) . 150
Displacement, mm .

100
897
50

0 892

-50
887
-100

-150
882
-200 Refer Figure C4.5a for legend

-250 877
70 72 74 76 78 80 82 84 86 88 90
Time (years since end of construction)
Figure 7.158: Belle Fourche dam closure section; post construction (a) settlement and
(b) displacement normal to dam axis of the crest and slopes over the period 1985 to
2000 (73 to 89 years post construction).

7.11.2.2 San Luis dam (Section 2.2 of Appendix G)

San Luis dam is a zoned earth and rockfill embankment with very broad central earthfill
zone constructed in the 1960’s. The embankment is of 116 m maximum height and
5650 m crest length. Foundation conditions vary along the length of the embankment,
from deep alluvial deposits in the floodplain area to bedrock at shallow depth on
hillslopes. Changes to the embankment design were made according to the foundation
conditions (refer Section 2.2 of Appendix G). A section at the slide area is shown in
Figure 7.159.
The very broad central earthfill core was constructed using mainly low to medium
plasticity sandy clays, sourced from alluvial terrace and floodplain deposits, compacted
by tamping rollers in 150 mm layers to high density ratio at moisture contents on
average 1.2% dry of Standard optimum. The miscellaneous fill zone (Zone 3) in the
outer downstream shoulder and upstream toe regions comprised materials ranging from
Zone 1 type earthfills to weathered rock, and were placed and compacted in 300 mm
layers.
Full supply level was reached in July 1969, 2.1 years after construction, and post
first filling the reservoir has been subjected to a seasonal drawdown ranging in
magnitude from 5 to about 65 m (Figure 7.162a).
Chapter 7: Deformation behaviour of embankment dams Page 7.266

Figure 7.159: San Luis dam, section of the slide in the upstream slope in 1981 at Station
135+00 (courtesy of United States Bureau of Reclamation).

In September 1981, some 14 years after construction, a slide occurred in the upstream
slope during large drawdown (Von Thun 1988; Stark and Duncan 1987, 1991). The
deep-seated slide (Figure 7.159) was located on the left abutment in a region where the
original ground surface under the embankment sloped in an upstream direction at about
10 to 15 degrees. It was approximately 460 metres in width (from Station 122 to 137),
1 million cubic metres in volume, and slid a distance of about 20 m.
Von Thun (1988) considered that the persistent, but minor, longitudinal crest
cracking over the 14 years up to the slide and settlement behaviour of SMPs on the
upstream at Stations 136 and 138 as indicators of “abnormal” deformation behaviour
prior to the slide and precursory warning signs of potential instability. Although, he
added that, actual prediction of the timing of the slide was not possible from the
monitored deformation.
Von Thun (1988) suggests that in the area of the slide the crest cracking “was
associated with the saturation and progressive straining of the slopewash on the
hillsides”. But, the longitudinal crest cracking was not confined to the region of the
slide, it was more of a general occurrence along the length of the embankment. This is
confirmed by comparison of the displacements along the line of SMPs upstream of the
dam axis with those downstream of the dam axis, which showed that leading up to the
slide lateral spreading of 50 to 150 mm occurred along most of the embankment length.
In addition, most of the differential displacement in the region of the slide occurred
during first filling (refer Section 2.2 of Appendix G). However, there is some query
over the “original zero” and the accuracy of the lateral displacement measurements in
the 1960’s and 1970’s.
Chapter 7: Deformation behaviour of embankment dams Page 7.267

For the overall embankment it could be said that the longitudinal crest cracking is
likely to be associated with differential deformations due to progressive development of
the phreatic surface in the dry placed and brittle earthfill, and the stress conditions
developed in the earthfill. Therefore, it is difficult to consider the crest cracking as an
indicator of “abnormal” deformation behaviour and a precursor to slope instability in
this case, unless persistence of the cracking was confined to the vicinity of the slide.
With respect to the settlement behaviour of SMPs on the upstream slope, Von Thun
(1988) comments that the very large differential between the actual and predicted post
construction settlements at Stations 136 and 138 (Figure 7.160) was a clear indicator of
the “anomalous” behaviour in the region of the slide prior to the failure. The SMPs on
the upstream slope at Stations 136 and 138 are located several metres above and
therefore outside of the initial slide area as shown in Figure 7.159.
Figure 7.161 presents the actual post construction settlements measured to the date
of the last reading prior to the slide for SMPs at 13 m upstream and those at 6.5
downstream of the dam axis. The figure shows that at most locations the measured
settlements up and downstream of the crest are similar, however, a differential is
notable for the SMPs from Station 130 to 140, in particular at about Station 136 where
the slide initiated. In addition, the settlement versus time plot of SMPs in the region of
the slide (Figure 7.162) shows that during the first large drawdown period from 8 to 10
years after construction an acceleration in settlement rate was measured for the SMPs
on the upstream slope. When the settlement is normalised with respect to the height
from the SMP to foundation level (Figure 7.162b) the magnitude of the settlement
during drawdown was greater for the SMP at Station 136. This trend in the settlement
behaviour might be considered “abnormal” in comparison to other SMPs on the
upstream slope at San Luis dam both in the slide vicinity and elsewhere along the
embankment.
In summary, the settlement behaviour of the SMP on the upstream slope at Station
136, located in the region of but not within the slide area, during the first large
drawdown appears “abnormal” compared to similar embankments and more importantly
to other SMPs on the upstream slope at San Luis dam. During the next large drawdown
in 1981, at 14 years after construction, the slide in the upstream slope occurred. In
hindsight, the settlement behaviour of this SMP could be considered as a precursory
sign of the failure, although, it would be almost impossible to predict the occurrence of
the slide based on the deformation records as concluded by Von Thun (1988).
Chapter 7: Deformation behaviour of embankment dams Page 7.268

The longitudinal crest cracking at San Luis dam is not considered as a reliable
indicator of “abnormal” deformation behaviour and precursory sign of slope instability
in this case because of its widespread occurrence along the embankment.

Figure 7.160: San Luis dam, difference between actual and predicted settlement (Von
Thun 1988)

-0.3%
SMPs at 6.5 m downstream (1/2/1981)
.

-0.2%
(% height from SMP to foundation)

SMPs at 13 m upstream (11/6/1981)


-0.1%

approx.
0.0%
width of slide
Settlement

0.1%

0.2%

0.3%

0.4%

0.5% Note: from Stn 55 to 102 settlement divided by


depth from SMP to bedrock foundation.
0.6%
0 20 40 60 80 100 120 140 160 180
Station

Figure 7.161: San Luis dam, comparison of settlement between SMPs at 13 m upstream
of dam axis and SMPs at 6.5 m downstream.
Chapter 7: Deformation behaviour of embankment dams Page 7.269

0 166
a)

Reservoir Level (elevation, m) .


50 141
Settlement (mm) .

100 116

150 Stn 138 - 6.5m downstream 91


Stn 136 - 13m upstream
Stn 128 - 13m upstream
Stn 144 - 13m upstream
200 66
Reservoir

250 41

time of slide
300 16
0 5 10 15 20 25 30 35
Time (years since end of construction)

0.0%
b)
(% height from SMP to foundation) .

0.1%

time of slide
0.2%
Settlement

0.3%

0.4%

0.5% accelation in settlement


rate during previous Stn 138 - 6.5m downstream
larger drawdown events Stn 136 - 13m upstream
0.6% Stn 128 - 13m upstream
Stn 144 - 13m upstream

0.7%
1 10 100
Time (years since end of construction)
Figure 7.162: San Luis dam, post construction settlement of SMPs in the vicinity of the
slide area.

7.11.3 THE EFFECT OF THE DEVELOPMENT OF THE PHREATIC SURFACE ON THE


DISPLACEMENT OF THE CREST AND DOWNSTREAM SHOULDER.

For a number of the earthfill embankments or zoned embankments with very broad core
width the displacement of the crest and/or downstream shoulder shows an increase in
the rate of displacement long-term (Figure 7.76 for the crest, and Figure 7.80 for the
downstream shoulder). The affected case studies include Horsetooth, Soldier Canyon,
Chapter 7: Deformation behaviour of embankment dams Page 7.270

Spring Canyon, Dixon Canyon, Pueblo (both left and right abutment embankments),
San Luis and Belle Fourche dams.
For a number of embankments of very broad core width constructed of low
permeability earthfills, the phreatic surface developed gradually within the embankment
over tens of years (refer Section 7.6.3 item f). Changes in the reservoir operation are
also likely to influence the phreatic surface. Maintaining the reservoir at higher
elevations for longer periods that previously experienced will result in a gradual rise in
the phreatic surface. The effect of a slow increase in the phreatic surface will cause a
gradual change in the effective stress conditions within the embankment and softening
of the material strength properties as the degree of saturation increases, particularly if
the core is yielded or dilated. These changes are likely to influence the deformation
behaviour and are considered a factor in the unusual observation of higher rates of
displacement post first filling.
At Belle Fourche dam the increase in displacement rate of SMPs on the crest and
downstream slope is observed after 82 years, and is approximately coincident with a
period of higher average reservoir level (Figure 7.158). In addition, the increase in
settlement rate of the downstream edge of the crest and downstream slope after year 85
may also reflect a change in effective stress conditions due to a rising phreatic surface
under the high average reservoir level.
At all the Horsetooth Reservoir embankments (refer Section 3.3 in Appendix G), an
increase in displacement rate to downstream was measured many years after the end of
first filling for SMPs on the crest and downstream slope. USBR (1997) indicate that the
pore water pressures within the earthfill zones of all embankments were still rising more
than 50 years after construction. The increase in pore water pressures suggests that
equilibrium conditions are yet to be reached, indicating that progressive wetting up is
still occurring.

7.11.4 OTHER CASES OF POTENTIAL “ABNORMAL” DEFORMATION BEHAVIOUR

7.11.4.1 Horsetooth dam


As previously discussed, the foundation had a significant influence on the post
construction deformation behaviour at Horsetooth dam. The embankment design over
the broad gully region (Figure 7.163) was such that the bulk of the embankment was
Chapter 7: Deformation behaviour of embankment dams Page 7.271

founded on the over-burden soils and the cut-off trench to bedrock was located
upstream of the dam axis. The depth of over-burden soils in the broad gully region was
up to 10 to 20 m.
Settlements in the foundation under the centreline of the embankment were very
large, estimated at almost 15% of the depth of the over-burden (500 to 1000 mm at the
IVM locations). Post construction, the settlement of the foundation was also very large
(325 mm at IVM A or up to 5% of the depth of the over-burden over nearly 50 years),
most of which (about 70%) occurred on first filling. The post construction vertical
strains in the foundation were very much larger than in the earthfill, where they average
1.05% at 45 years after construction.
The large differential settlement between the foundation and earthfill in the cut-off
had the effect of causing a downstream rotation of the embankment downstream of the
cut-off trench. As shown in Figure 7.164, the upper upstream shoulder to downstream
shoulder displaced downstream during the period of first filling from 2 to 4.5 years after
construction.
Compared to similar embankments the displacement at Horsetooth dam on first
filling is an outlier (Figure 7.38 to Figure 7.40), but not “abnormal” because of the
influence of the compressible foundation.

Figure 7.163: Main section at Horsetooth dam (courtesy of the United States Bureau of
Reclamation).
Chapter 7: Deformation behaviour of embankment dams Page 7.272

300
end of first filling

250
(downstream is positive) .
Displacement (mm) .

200

150

100

50 Crest, downstream edge (Stn 10)


19m upstream (Stn 10)
51m downstream (Stn 10)
0

-50
1 10 100
Time (years since end of construction) Note: SMP offsets from centre of crest

Figure 7.164: Horsetooth dam, post construction displacement of SMPs near to the main
section.

7.11.4.2 Pueblo dam, left abutment embankment (Section 3.4 of Appendix G)


The left embankment at Pueblo dam (Figure 7.165) is of 37 m maximum height and
about 1100 m crest length. The main earthfill zone consisted of clay to gravel size
alluvium placed in 150 mm layers and compacted by tamping rollers. The moisture
specification called for the mean moisture content to be in the range 0.5% dry to 1.5%
dry of Standard Proctor optimum moisture content. Coarser and more permeable soils
were used in the outer up and downstream regions of the main earthfill zone.
During construction (in the early to mid 1970’s) shear type deformations were
measured in the downstream toe region of the left embankment when the embankment
was within about 7 m of design crest level. Inclinometers showed that the deformations
were concentrated along weak bentonitic clay seams in the foundation. By end of
construction total shear type displacements were estimated (by the USBR) at more than
150 mm at the downstream toe of the left embankment. The shear deformations
virtually ceased shortly after the end of construction. In the early 1980’s a stability
berm was constructed along the downstream toe of the left abutment embankment due
to concerns over the potential limiting stability of the downstream slope.
Chapter 7: Deformation behaviour of embankment dams Page 7.273

Figure 7.165: Pueblo dam; cross section of left abutment embankment at Station 75

Post construction settlements of the downstream shoulder of the left abutment


embankment at Pueblo dam were large (Figure 7.166), and in comparison to similar
type embankments the magnitude is “abnormally” high (Figure 7.50 and Figure F2.10a
of Appendix F). The long-term settlement rate of the downstream shoulder is on the
high side, but not “abnormally” so (Figure 7.78).
In the vicinity of Station 75 borehole records indicate the soils are mainly weathered
shales, for which the post construction deformation is likely to be limited and therefore
not greatly influence the deformation behaviour of the embankment. In addition, the
shear type deformations observed in the foundation during construction had virtually
ceased at the end of construction, and therefore would not influence the post
construction deformation behaviour of the downstream shoulder with the stability berm
added.
As Figure 7.166 shows, the first settlement reading after the base survey of the SMP
on the downstream slope is very high (0.75% or almost 150 mm) and contributes
significantly to the large post construction settlement (42% of the total settlement
measured over 23 years). After this first reading the settlement behaviour is similar in
magnitude to other SMPs on the embankment crest and upstream slope, although the
settlement rate is slightly higher. Other SMPs along the downstream slope of the left
embankment abutment show similar settlement behaviour to that at Station 75.
It is not clear what the cause of this observed settlement behaviour for the
downstream slope is. The influence of construction of the stability berm is not the cause
because construction started in 1980 some 5 to 5.5 years after end of construction, after
the period of large settlement. It is possible that it could be survey error. Apart from
the settlement of the downstream shoulder, the post construction deformation behaviour
Chapter 7: Deformation behaviour of embankment dams Page 7.274

at the left abutment embankment of Pueblo dam is “normal”. Further details on the
deformation behaviour at Pueblo dam are discussed in Section 3.4 of Appendix G.

2.0% 1490

Reservoir Level (elevation, m) .


1.6% 1480
.
Settlement (mm)

1.2% 1470

0.8% 1460

32m upstream (Stn 77.5)


0.4% Crest, upstream edge (Stn 75) 1450
Crest, downstream edge (Stn 77.5)
40m downstream (Stn 75)
Reservoir
0.0% 1440
0 5 10 15 20 25

Time (years since end of construction)

Figure 7.166: Pueblo dam, post construction settlement of the left embankment near
Station 75.

7.12 “ABNORMAL” D EFORMATION BEHAVIOUR OF P UDDLE CORE


EARTHFILL DAMS
For puddle core earthfill dams, trends of “abnormal” deformation behaviour are specific
to the long-term deformation behaviour given the period of construction of these
embankment types. Identification of “abnormal” deformation behaviour is from
comparison with the deformation behaviour of other puddle dams, mainly in terms of
rates of deformation, as well as the deformation trend in relation to the creep model
under constant stress conditions. The limited number of records from the end of
construction does not allow for comparison based on the magnitude of deformation.
As previously discussed, what may initially be termed “abnormal” may later be
proven by investigation and additional monitoring to in-fact be “normal”. In the case of
puddle core earthfill embankments, the large deformations observed during historically
large drawdown events (Section 7.7.3.5) could well be considered as “abnormal” when
compared to the deformation trends under normal reservoir operating conditions.
Chapter 7: Deformation behaviour of embankment dams Page 7.275

7.12.1 COMPARISON WITH THE DEFORMATION BEHAVIOUR OF OTHER PUDDLE DAMS

Most of the available data to assess the long-term deformation behaviour of a puddle
core earthfill embankment relates to crest settlement (Table 7.24, and Table F1.5 in
Section 1 of Appendix F). There is some, but limited, data on settlement of the
shoulders and at the downstream toe, and on horizontal displacement.
In terms of crest settlement, comparisons can be made as follows:
• Total settlement from the end of construction (Figure 7.92). Although, only limited
data is available for comparison. The age of the embankment, embankment height,
compaction methods and moisture control affect the total settlement. From Figure
7.92:
− For dams constructed after about 1900, and particularly after about 1930, the
total settlement at about 30 to 50 years is less than 4% of the total height of the
embankment. In comparison, the total crest settlement of Hollowell Dam
(completed in 1937) stands out as being “abnormal” and is discussed later.
− For older puddle dams, there is insufficient information to make a judgement on
what could be expected as “normal” total settlement. Yan Yean and Hope
Valley dams show total vertical settlements of 8 to 13% of the embankment
height at more than 100 years.
• The long-term crest settlement rate, SLT, under normal reservoir operating
conditions. As discussed in Section 7.7.3, long-term rates of crest settlement depend
on a number of factors, in particular the level of fluctuation of the reservoir and the
pore water pressure response in the earthfill zone upstream of the puddle core.
− For “steady state” conditions SLT is typically less than 1% per log cycle of time
(settlement as a percentage of the embankment height). “Steady state”
conditions are defined as virtually steady reservoir operating levels or, in the
case of fluctuating reservoir level, negligible to minor changes in pore water
pressure in the earthfill zone upstream of the puddle core to fluctuations in the
reservoir level.
− For embankments subject to reservoir fluctuation and in which the upstream
earthfill is permeable (i.e. the changes in pore water pressure in the earthfill zone
upstream of the puddle core approximate the fluctuations in reservoir level), SLT
rates of 2.6% to 7.4% per log cycle of time have been measured, and are
considered to be “normal”.
Chapter 7: Deformation behaviour of embankment dams Page 7.276

− Figure 7.98 presents data on crest settlement versus drawdown height for dams
subject to reservoir fluctuations greater than 20% of the dam height and in which
the earthfill zone upstream of the puddle core is permeable. The data has been
adapted from Tedd et al (1997b). This figure shows some correlation between
height of drawdown and permanent crest settlement, and could be used as
preliminary tool for assessment of “abnormal” deformation. However, the use
of Figure 7.98 is limited given that the data set comprises older UK dams (pre
1910) of similar height and dimensions, and constructed using similar materials.

Other than for crest settlement, the deformation data on puddle core earthfill dams is
limited due to the limited data available. Comparisons based on settlement of the
embankment shoulders are not possible. However, assessment of the horizontal
displacement of the crest and downstream slope is possible.
The post construction deformation behaviour of earthfill and earth-rockfill dams
indicates that the general displacement trend of the crest and downstream shoulder is for
downstream displacement at a decreasing rate (on log time scale) with time, eventually
reaching a point where the net displacement is negligible with some fluctuation about
the general trend due to fluctuations in reservoir level. A similar analogy is possible for
puddle dams. The limited amount of available data tentatively suggests:
• Negligible long-term rates of downstream displacement of the crest and downstream
slope for case studies at “steady state” conditions.
• Long-term displacement rates of the downstream crest and slope of up to 3 to 5
mm/year downstream for embankments where the upstream shoulder is relatively
permeable and the reservoir is subject to relatively large drawdowns (more than
about 20% of the embankment height) and the fluctuation is within the range of
normal reservoir operation.

The deformation behaviour of the following case studies is considered to be


“abnormal” based on comparison with other puddle dams:

(a) Hollowell dam


The crest settlement of Hollowell dam in the period shortly after embankment
construction. The magnitude of total crest settlement of Hollowell dam in comparison
to puddle dams of similar age is significantly greater (Figure 7.92). At 10 years after
Chapter 7: Deformation behaviour of embankment dams Page 7.277

the end of construction crest settlement at Hollowell dam was more than 6% compared
to less than 2% for Selset and Burnhope dams.
Kennard (1955) indicates the large deformations at Hollowell dam were due to the
marginal factor of safety of the downstream slope at the end of construction as a result
of the high pore water pressures in the foundation. Localised shear type deformations
occurred during construction and measures were undertaken to improve the factor of
safety. The long-term settlement rate for Hollowell dam more than 10 years after the
end of construction is about 0.7% per log cycle of time, which is comparable to dams of
similar characteristics.

(b) Yan Yean dam (Section 5.2 of Appendix G)


The high long-term rate of settlement of the downstream edge of the crest at Yan Yean
dam measured more than 130 years after the end of construction is considered
“abnormal”.
Yan Yean dam has been categorised in the class of “steady state” embankments
based on the minor pore water pressure response in the earthfill zone upstream of the
puddle core compared to the change in reservoir level. SLT rates for the crest would be
expected to be less than about 1% per log cycle of time, which is the case for the
upstream edge of the crest.
However, for the downstream edge of the crest SLT is a maximum of 11% over a
period of 14 years from 128 to 142 years after construction, much greater than would be
expected of “normal” type behaviour. High rates of settlement and displacement of the
downstream crest and slope were also observed over this period of monitoring (up to 8
to 10 mm/year settlement, and up to 4.5 - 5.5 mm/year downstream crest displacement).
These high rates are also considered indicative of “abnormal” deformation behaviour.
The increasing rate of deformation with time of several SMPs (refer Section 5.2 of
Appendix G) is also a strong indicator of “abnormal” deformation behaviour.
The deformation behaviour of the downstream slope is considered to be independent
of the reservoir level operation and possibly indicative of a tertiary creep (or creep to
failure) phase of movement (refer Section 7.12.2). Investigations showed fissured clays
in the foundation were contributing to a marginal stability condition, and a berm has
now been constructed to improve stability under normal operating conditions and
earthquake, and internal erosion and piping control.
Chapter 7: Deformation behaviour of embankment dams Page 7.278

(c) Hope Valley dam (Section 5.3 of Appendix G)


For Hope Valley dam the rate of downstream displacement of the upper downstream
slope of 5 mm/year over the monitoring period from 118 to 128 years after construction
maybe on the high side, but would not be classified as “abnormal” from the available
data. But, the deformation behaviour could be indicative of marginal stability condition
of the downstream slope. Hope Valley dam has since been remediated and a
downstream stabilising berm added (Gosden et al 2002).

7.12.2 GENERAL MOVEMENT TRENDS INDICATIVE OF DEFORMATION TO FAILURE

From the basic creep model of time dependent deformation under constant deviatoric
stress conditions (Singh and Mitchell 1968; Mitchell 1993), an increasing rate of
deformation under constant stress is indicative of a deformation to a failure condition
(i.e. a tertiary creep phase of deformation). A localised trend of increasing rate of
deformation at one SMP may indicate shallow surficial deformations or a faulty SMP,
and not a marginal stability condition and potential failure condition. It is therefore
important that a similar deformation trend is evidenced by other SMPs on the
embankment in the close proximity to each other and that a vector plot of the
deformation at a section is indicative of a potential deep-seated movement.
The deformation trends of several SMPs on the downstream crest and slope of Yan
Yean dam (Figure 7.167) show an increasing rate of deformation, indicating a potential
tertiary creep mode of deformation, that is considered to be independent of reservoir
level fluctuation. The possible mechanisms involved in the deformation behaviour of
the downstream slope are considered to be a combination of stress changes at the top of
the slope due to moisture infiltration into cracks when the reservoir is in a drawdown
condition and to progressive failure of the foundation or saturated base of the
embankment. It would appear that stick-slip type deformation behaviour of the
downstream toe at several locations indicates that the movements are being driven from
the top of the slope resulting in build up of stresses in the foundation (or saturated base
of the embankment). Continued movements of this type within the over-consolidated
foundation could result in strain weakening and a progressive failure. As previously
mentioned a downstream berm has now been constructed to improve stability.
Chapter 7: Deformation behaviour of embankment dams Page 7.279

25
SMP 1501 - crest, upstream edge
Possible tertiary creep phase for
SMP 1502 - crest, downstream edge SMPs 1502, 1503 and 7502
Settlement Rate (mm/year) . 20 SMP 1503 - downstream slope
SMP 7502 - crest, downstream edge

15

10

-5
126 128 130 132 134 136 138 140 142
Time (years since end of construction)

Figure 7.167: Yan Yean dam; settlement rate versus time of SMPs at Chainage 150 and
750 m.

7.12.3 OTHER INDICATORS OF “ABNORMAL” DEFORMATION BEHAVIOUR

Several other examples from the long-term monitoring trends of puddle core earthfill
embankment case studies are considered to be indicative of possible “abnormal”
deformation behaviour. These were observed at Ramsden and Hope Valley dams, both
of which are discussed further in Section 5 of Appendix G.
For Ramsden Dam, the abnormal deformation behaviour is related to the internal
deformation of the puddle core on the abnormally large drawdown in 1988 to 1989. It
is considered possible that shearing occurred in the puddle core between 10 m and 11 m
depth below the crest as indicated by the localised high vertical strain and possible
upstream shear displacement at this depth. This is associated with a permanent
settlement of 51 mm at the crest, smaller permanent settlements of the upstream slope (2
to 10 mm) and 7 mm permanent upstream displacement of the crest.
For Hope Valley dam (Appendix G, Section 5.3) the displacement rate of the upper
downstream slope (5mm/year) is considered to be relatively high in comparison to other
puddle dams. The rate is similar to that of Yan Yean dam, although the rate at Hope
Valley dam was virtually constant and not increasing. In comparison to the crest of
Walshaw Dean and Ramsden, which also shows a steady downstream displacement
rate, the rate is higher at Hope Valley dam, however this may be due to type of material
Chapter 7: Deformation behaviour of embankment dams Page 7.280

used as shoulder filling and possibly a lower factor of safety of the downstream
shoulder.
Based on the available deformation records for Hope Valley dam and in comparison
with other puddle dams, it is considered that a significant proportion of the movement
of the downstream slope can be attributed to long-term creep and cyclic stress changes
associated with reservoir level fluctuations. However, it is not possible to conclude that
these are solely the explanation for the movement without considering the possibility
that the downstream shoulder is in a marginal stability condition. The discerning factor
is the high and constant rate of downstream displacement. As previously mentioned a
downstream berm has now been constructed to improve stability.

7.13 SUMMARY OF “ABNORMAL” D EFORMATION BEHAVIOUR


The purpose of identification of “abnormal” deformation behaviour of embankment
dams is for the early detection of potential problems, mainly with respect to slope
instability. But, it may also have implications with respect to internal erosion and
piping because the deformations may lead to cracking and softening. Following
identification of “abnormal” deformation behaviour it is necessary to understand the
mechanism/s causing the observed behaviour, which may incorporate investigation and
analysis, and may eventually lead to some form remedial works. In most cases
“abnormal” deformation behaviour does not equate with stability issues for the
embankment, but in a number of cases it has.
The methods of identification of “abnormal” deformation behaviour from the case
study analysis are mainly as outliers to the “normal” deformation behaviour of similar
embankment types in terms of magnitude, rate and trends. In summary, the methods
are:
• During construction (for earthfill and zoned earth and earth-rockfill embankments),
outliers in terms of:
− Total core settlement during construction (Figure 7.23, Table 7.14 and Table
7.15).
− Magnitude of vertical strain within the core at a given depth or effective vertical
stress level (Figure 7.18, Figure 7.20 and Figure 7.21).
• Post construction, outliers in terms of:
− Magnitude of settlement or displacement.
Chapter 7: Deformation behaviour of embankment dams Page 7.281

− Settlement rate (rate in terms of log time) including:


• Magnitude of the settlement rate
• Acceleration in rate over short periods of time, often, but not always, post
first filling on drawdown.
− Displacement:
• Direction of displacement
• Magnitude of the long-term displacement rate (rate in terms of log time)
• Change in direction of the underlying general trend, i.e. the trend outside of
that due to reservoir fluctuation.
• Short periods of non-recoverable displacement post first filling, such as a
permanent upstream displacement on large drawdown.
• Development of localised regions of high strain (during or post construction)
indicative of the formation of a shear surface, and the ongoing deformation within
these regions. The case study analysis concentrated on vertical strains in the core
region of the embankment, but this would be equally applicable to concentration of
lateral strain in the core or foundation measured in inclinometers.
• The tertiary creep analogy from the model of creep under constant stress conditions.
Tertiary creep (or creep to failure) is creep at an increasing rate (rate in terms of
normal time) with time and is an indication of the onset of failure. Primary creep, or
creep at a decreasing rate (rate in terms of normal time) with time is indicative of
“normal” type behaviour.

Several important aspects on the deformation behaviour should be noted:


• For zoned earth and rockfill dams, “abnormal” or potentially “abnormal”
deformation behaviour was much more likely in embankments where the rockfill
was susceptible to large settlements due to collapse compression on wetting or was
of high compressibility. These include rockfills that are poorly compacted, dry
placed and poorly to reasonably compacted, dumped and sluiced rockfills,
weathered rockfills, and rockfills of rock type susceptible to large loss in unconfined
compressive strength on wetting. Conversely, sound rockfills that are wetted and
well and reasonably to well compacted are, in most cases, not susceptible to large
collapse compression on first filling and the overall post construction deformation
behaviour of the embankment is “normal” and generally of limited magnitude.
Chapter 7: Deformation behaviour of embankment dams Page 7.282

• The development of a shear zone in the earthfill core does not equate with a
marginal factor of safety, although it may. The shear zone may be a result of
differential settlement between the core and shoulder or a lack of support from the
shoulders. The core types within which shear surfaces developed (and the timing of
the shear development) included:
− Compacted silty sands to silty gravels. Shear surfaces in the core generally
developed during first filling or with further movements shortly thereafter (often
drawdown related). High shear stresses at the interface between the core and
upstream rockfill shoulder (that developed due to the greater settlement of the
upstream shoulder as it collapse compressed on wetting) were considered the
mechanism in most cases for the shear development.
− Dry placed clayey sand earthfills. Shear surfaces in the core generally
developed during first filling with further deformations associated with reservoir
operation.
− Dry placed sandy clay and clay earthfills. Shear surface developed on first
filling (e.g. Ataturk dam) or many years post first filling, usually on large
drawdown (see below).
− Wet placed clayey earthfills. Examples of shear surface development during
construction and on first filling, as well as post first filling.
• Longitudinal cracking does not equate with a marginal factor of safety. In most
cases it may simply be due to differential deformation between the zones in the
embankment and may only occur during the period of and shortly after first filling.
However, persistent longitudinal cracking may be indicative of a marginal stability
condition (refer below).
• Acceleration of the deformation of the crest and upstream slope on large drawdown
and resultant permanent deformations do not equate with marginal stability, but they
may. These type of deformations are not uncommon on historically large
drawdowns where yielding may occur under effective stress levels not previously
experienced (such as observed in several puddle dams). Persistent observations of
acceleration of the deformation on large drawdown may be indicative of a marginal
stability condition (refer below).
• For several central core earth and rockfill embankments with dry placed clay cores
and dry placed and/or poorly compacted rockfills, shear surfaces developed (or were
Chapter 7: Deformation behaviour of embankment dams Page 7.283

thought to have potentially developed) in the central region of the core many years
after the end of construction. In the case of Eppalock dam it was considered that
significant cracking in the core preceded the shear development as indicated by
observed cracking and/or the change in displacement trend of SMPs on the crest.
The cracking prior to shear development was considered a significant factor in the
development of the shear surface, which was first identified during a large
drawdown (not necessarily the first large drawdown). This type of shear surface
development was considered to have possibly developed at several other similar
type embankments (i.e. dry placed clayey core with rockfill shoulders susceptible to
collapse compression on wetting).

The issue of potential instability or marginal stability of the embankment is the


foremost aspect of any deformation monitoring that is undertaken given the potentially
catastrophic consequences of a slope failure condition in an embankment dam. Some
guidelines from the analysis of those case studies where failure occurred or where the
embankment was considered to be in a marginal stability condition are:
• Instability during construction:
− Deformation behaviour during shutdown periods. Large and ongoing
deformations during shutdown or acceleration in the rate of deformation (rate in
terms of normal time) may indicate marginal stability or an impending failure
condition.
− The incremental magnitude of deformation with increasing dam height or stress
level (refer Penman 1986). High or increasing (e.g. Carsington dam)
incremental magnitudes of deformation may be indicative of a marginal stability
condition or the onset to a failure condition.
− Localised regions of “abnormally” high strain as measured in internal settlement
gauges or inclinometers may be an indicator of internal shear and potentially a
progressive failure mechanism. Increases in the incremental magnitude of
localised strain may be indicative of the onset to a failure condition.
• Post construction on drawdown:
− Persistent development of cracking on large drawdown
− Persistent acceleration in settlement and/or displacement of SMPs on the crest
and/or upstream shoulder on large drawdown.
Chapter 7: Deformation behaviour of embankment dams Page 7.284

− A similar or increasing magnitude of non-recoverable deformation on


consecutive large drawdowns of similar magnitude (consecutive meaning 1, 5,
10 or more years between drawdowns of similar magnitude interspersed with
smaller magnitude drawdowns).
− Acceleration in settlement or displacement that is confined to one region of the
embankment, such as observed at San Luis dam in the region of the slide in the
large drawdown preceding the one that triggered the slide (refer Section
7.11.2.2).
− Localised regions of high magnitude deformation compared to similar locations
elsewhere on the embankment or compared to predicted deformations (Von
Thun (1988), for San Luis dam).
• Post construction, downstream shoulder:
− Tertiary creep phase of deformation (e.g. Yan Yean dam).

From the study of failures in embankment dams (excludes failures in hydraulic fill
dams) in Chapter 5, the conversion of a significant portion of the potential energy into
kinetic energy at or post failure is required for acceleration of the slide mass and large
post failure deformation. Factors contributing to this conversion in energy for the case
studies of slides in embankment dams included:
• Potential for material strain weakening on shearing at or after failure in drained or
undrained loading conditions (including static liquefaction of structured soils in
undrained loading).
• Brittleness in the slide mechanics, such as due to internal brittleness or toe
buttressing, and/or brittleness on the lateral margins.
• The slope failure geometry and orientation of the surface of rupture.

The deformation behaviour has important implications on the embankment


performance related to the potential for internal erosion and piping, as well as the long-
term stability of the embankment. With respect to internal erosion and piping a number
of aspects are considered to increase the potential for formation of a seepage path
through the core of a zoned embankment, including:
• Cracking. The influence of cracks across the core is readily apparent. The possible
causes of cracking associated with deformation behaviour are discussed by others
Chapter 7: Deformation behaviour of embankment dams Page 7.285

(Sherard 1973; amongst others). In addition to these mechanisms of crack


formation, localised large deformations are likely to give 3D cracking that may
persist through the core. Also, further differential deformations post construction,
such as between the core and shoulders or core and foundation, can open up or
widen existing cracks.
• Further differential deformations post construction, such as between the core and
shoulders or core and foundation, can result in further arching and reduction in the
stress conditions within the core, thereby increasing the potential for hydraulic
fracture.
• Shear or softened zones that persists across the width of the core are also potential
seepage paths.

It is evident from the deformation behaviour at a number of embankments that the


softening process in ongoing and can lead to deterioration of the embankment over time,
and potentially a gradual reduction in the factor of safety of the embankment’s stability.
The clearest example is evidenced by the timing of upstream slope instability on
drawdown in rolled earthfill embankments (e.g. Belle Fourche and San Luis dams)
where the embankment was subjected to several large drawdown events before the
failure occurred. Part of the mechanism associated with these and other upstream slope
failures is considered to be progressive strain weakening in undrained loading under
cyclic reservoir operation within the well-compacted over-consolidated rolled earthfill
shoulder. This mechanism was recognised by Stark and Duncan (1987, 1991) as
significant in the failure at San Luis dam and is reflected in the composite shear strength
concept recommended by Duncan et al (1990) for limit equilibrium analysis under
drawdown. Other softening processes associated with the deformation behaviour of
embankment dams are:
• Lateral spreading and/or cracking of the core. Lateral spreading (and therefore
cracking) in central core earth and rockfill dams is more significant where the
rockfill is susceptible to large settlements from collapse compression on wetting.
Part of the reason for this is considered to be due to the transfer in stress to the core
associated with the greater settlement of the shoulders, and due to the reduction in
support that is provided by shoulders that have suffered large deformations.
• In earthfills susceptible to collapse compression on wetting, the reduction in shear
strength and compressibility properties is significant. These earthfills are virtually
Chapter 7: Deformation behaviour of embankment dams Page 7.286

normally consolidated on softening due to wetting or post collapse. In the case of


Hume dam (Cooper et al 1997) a significant factor in the marginal stability of the
downstream shoulder of the concrete core-wall earthfill embankment at near full
supply level was the low undrained shear strength of the near normally consolidated
saturated earthfill in the lower portion of the downstream slope.

In limit equilibrium analysis therefore, consideration should be given to the potential


for softening and the influence of the embankment deformation behaviour on the shear
strength properties of the embankment materials. Some guidelines are:
• Use the method by Duncan et al (1990) for analysis of the upstream shoulder under
drawdown. It takes into consideration the effects of progressive strain weakening of
over-consolidated earthfills under cyclic reservoir operation.
• For zoned earth and rockfill embankments where the shoulders are susceptible to
large settlements associated with collapse compression on wetting, the use of fully
softened c′ and φ′ parameters is wise for over-consolidated earthfill cores because of
the potential for softening from lateral spreading and cracking (in addition to the
softening on wetting from reservoir seepage).
• For earthfills susceptible to collapse compression (refer below) the collapsed
earthfill is likely to be near normally consolidated. The potential for development of
shear induced positive pore water pressures should be considered in the stability
analysis and this may require the use of undrained strength rather than a
conventional c′, φ′ analysis.
• Where a concentrated shear is developed residual strength parameters are
appropriate. Concentrated shears have developed within the core zone in a variety
of core types from silty sands/silty gravels to high plasticity clays, and in both over-
consolidated dry placed clays and wet placed clay cores of low undrained strength.
Consideration should also be given to the potential for strain localisation near to the
interface between zones (e.g. in the core near to interface with the shoulder, such as
observed at Ataturk dam (Cetin et al 2000)) and the use of residual strength
parameters for these shear zones.

It is widely recognised that poorly compacted and dry placed earthfills are
susceptible to collapse compression on wetting leading to large deformations (Gould
1954; Bernell 1958; Penman 1986; Charles 1997; amongst others). The susceptibility
Chapter 7: Deformation behaviour of embankment dams Page 7.287

of an earthfill to collapse compression is dependent on its material type, density and


moisture content at placement, and the level of stress within the embankment. The case
study evidence of earthfills suspected of collapse compression indicates:
• Dry placed and poorly compacted earthfills are susceptible to collapse compression
on wetting (e.g. outer earthfill zone in old puddle embankments (pre 1900/1920)).
There is no information from the database to suggest how dry poorly compacted
earthfills need to be placed for them to be susceptible to collapse compression.
• Formally compacted earthfills susceptible to collapse compression include:
− Earthfills placed on the dry side of optimum. Clayey earthfills placed drier than
about 2% dry of Standard optimum are susceptible. But, this will vary
depending on the material type, fines content, fines plasticity and compacted
density ratio. Silty sands and gravels and clayey earthfills with low fines content
or low plasticity fines may be susceptible when placed only 1% dry of Standard
optimum. Charles (1998) suggests the air voids be reduced to less than 5% in
clay earthfills to ensure the earthfill is not susceptible to collapse compression.
− Material types including silty sands and gravels, clayey sands and gravels and
sandy clays generally of low plasticity. Medium to high and high plasticity
clays do not appear to be susceptible to collapse compression (when well
compacted).
− Layer thickness and the variation in density within the layer are also important
considerations. The lower portion of thick placed layers, where the density is
often lower, is more susceptible to collapse compression than the upper, more
heavily compacted part of the layer.

7.14 SUMMARY AND M ETHODS FOR P REDICTION OF DEFORMATION OF


EMBANKMENT DAMS
This section presents a summary of the outcomes from analysis of the deformation
behaviour of earthfill, zoned earth and earth-rockfill, and puddle core earthfill
embankments from Sections 7.4 to 7.7 for use in prediction of, or comparison of the
deformation behaviour of an embankment. The methods for evaluation of the “normal”
deformation behaviour were developed to identify potentially “abnormal” behaviour,
Chapter 7: Deformation behaviour of embankment dams Page 7.288

and the figures and tabulated data provide a means for prediction or comparison to
similar embankment types.
In part, the summary is a pointer to figures and tables in this chapter relating to
specific aspects of embankment deformation behaviour. It also briefly summarises the
factors affecting the deformation behaviour as determined from the analysis as well as
bringing together some of the data within this chapter for ease of use.

7.14.1 EARTHFILL , AND ZONED EARTH AND EARTH-ROCKFILL DAMS

The methods for prediction of deformation behaviour of earthfill and zoned earth and
earth-rockfill dams are divided into two subsections, deformation during construction
and deformation post construction.

7.14.2 PREDICTION OF DEFORMATION BEHAVIOUR DURING CONSTRUCTION FOR


EARTHFILL AND ZONED EARTH AND EARTH-ROCKFILL EMBANKMENTS

Prediction of the deformations during construction of earthfill and zoned earth and
earth-rockfill dams are best undertaken by finite element methods. The difficulty with
these methods is in selection of properties and the constitutive model to represent the
various material zones, as well as consideration of the use of coupled models and
dealing with partial saturation of earthfills and the change in matric suction with stress
level.
The total and effective stress conditions established in an embankment during
construction are dependent on the embankment geometry, the embankment zoning
geometry, and the strength and compressibility properties of the embankment materials.
The simplest case to analyse is the elastic analysis of a homogeneous embankment on a
rigid foundation where non-linearity of material properties is taken into account in the
model used to represent the compressibility properties of the earthfill. Results show that
under the embankment centreline lateral strains are negligible and deformation is
predominantly in the vertical direction. The homogeneous model is shown to be a
reasonable assumption for the stress conditions under the embankment axis for:
• Zoned embankments with broad central earthfill zones for rolled earthfill placed dry
of Standard optimum (more than about 0.5% to 1% dry), where only limited to
negligible positive pore water pressures are developed during construction.
Chapter 7: Deformation behaviour of embankment dams Page 7.289

Analysis and the case study data suggests it is a reasonable assumption for
embankments with central core widths having a combined core slope (i.e. upstream
and downstream core slope combined) greater than about 1.5 to 2H to 1V, but it
may also be a reasonable simplification for combined central core widths down to
1H to 1V.
• Zoned embankments where the compressibility properties of the central earthfill
zone and gravelly or rockfill shoulders are similar; i.e., compacted sandy and
gravelly soils with non-plastic fines or low (less than about 20% finer than 75
micron) plastic fines contents, with shoulders of compacted gravels or rockfill. This
is on the proviso that pore water pressures generated in the core during construction
are small and potential plastic type core deformations due to lateral spreading of the
core are negligible.

Case study analysis shows that for dry placed earthfills the confined secant modulus
shows a gradual increase with increasing effective vertical stress (ignoring matric
suction). Therefore, reasonable numerical deformation solutions can be obtained
without consideration of pore water pressures provided the model used reasonably
approximates the compressibility of the earthfill. Values of confined secant modulus of
dry placed earthfills estimated from field data is presented in Table 7.13 and Figure 7.19
sorted based on material type. For dominantly sandy and gravelly earthfills the data is
further sorted based on fines plasticity and fines content. The data provides useful
bounds for comparison with laboratory test data on proposed core materials and
assistance in selection of the confined moduli for use in analysis.
At the other end of the spectrum is wet placed earthfill cores of low undrained shear
strength used in zoned earth and earth-rockfill embankments. Numerical analysis
modelling the core in undrained conditions is considered to provide a reasonable
approximation of the deformation behaviour for earthfills of low permeability, where
the deformation of the core occurs largely as undrained plastic type deformations. The
amount of lateral spreading of the core, which is influenced by the lateral stresses
developed in the core and the compressibility of the supporting shoulder zones, has a
significant influence the deformation of the core.
Modelling becomes more complex where:
• Pore water pressure dissipation in the core occurs during construction and the use of
a coupled model is desirable.
Chapter 7: Deformation behaviour of embankment dams Page 7.290

• Over the period of construction the initial deformation of the core is largely due to
compression of air voids, but in the latter stages largely occurs as plastic type
deformations in undrained conditions.

For rockfill and gravel earthfill zones, Chapter 6 provides information for use in
selection of the confined modulus properties based on intact rock strength, placement
method and particle size distribution.
Several methods have been developed for evaluation of the deformation behaviour
of embankment dams during construction that are useful for comparative purposes.
They are:
• Total core settlement for the period of construction. Estimation of the total core
settlement is based on embankment height, core material type and core width
(Figure 7.23, Table 7.14 and Table 7.15). This method was effectively used to
identify “abnormally” large deformations during construction for several
embankments.
• Vertical strain profile in the core at end of construction; Figure 7.18 for dry placed
earthfill cores, Figure 7.21 for wet placed dominantly sandy and gravelly cores
(SC/GC/SM/GM cores), and Figure 7.22 for wet placed clay cores. The figures
provide approximate bounds for “normal” type deformation behaviour. They can
also be used for evaluation of core deformation as construction proceeds. The
figures are useful for identification of regions of the core in embankments where
vertical strains were excessively large. Plots of vertical strain versus effective
vertical stress or height above the cross-arm interval (e.g. Figure 7.15 or Figure
7.20) for regions of high strain can then be used to assess the incremental vertical
strain and evaluate the possibility of shear type deformation.
• Lateral deformation of the core for central core earth and rockfill dams of thin to
medium core width (Section 7.4.2). The estimated lateral displacement ratio (LDR)
of the core at selected depths in a limited number of case studies is presented in
Figure 7.12 and Figure 7.13. The data indicates:
− LDR is influenced by the lateral stress developed in the core and so will be
greater for wet placed than dry placed cores, and earthfills of low permeability
(i.e. limited dissipation of pore water pressures during construction).
− The compressibility of the supporting shoulder zones has a significant influence
on LDR. LDR increases with increasing compressibility of the shoulders.
Chapter 7: Deformation behaviour of embankment dams Page 7.291

− The location of measurement affects LDR. Finite element analysis of a dam on


a rigid foundation showed LDR at end of construction to be a maximum at about
50 to 70% of the depth below crest level.

7.14.3 PREDICTION OF DEFORMATION BEHAVIOUR POST CONSTRUCTION FOR


EARTHFILL AND ZONED EARTH AND EARTH-ROCKFILL EMBANKMENTS

A number of figures and tables have been developed for evaluation and/or prediction of
the deformation behaviour post construction of SMPs on the crest and slopes of the
embankment. In summary they include:
• Settlement (based on zero time at the end of embankment construction):
− Magnitude of settlement for the crest and shoulders at 3, 10 and 20-25 years
after construction
− Settlement versus time plots for the crest and shoulders
− Settlement rate (rate in terms of log time)
• Lateral displacement:
− Of the crest, upstream shoulder and downstream shoulder on first filling (Section
7.5.3)
− Of the crest post first filling
− Lateral displacement versus time plots for the crest and shoulders (based on zero
time at the end of embankment construction)

(a) Deformation versus time plots


A large number of figures have been produced of deformation (i.e. settlement or
displacement) versus time, with zero time established at the end of embankment
construction. Table 7.26 and Table 7.27 are a pointer to the plots of settlement and
displacement versus time for the crest and shoulders respectively. Note that for zoned
earth and earth-rockfill dams with thin to thick core widths, the upstream edge of the
crest is included in the upstream shoulder region and the crest region is from the central
to downstream edge of the crest (refer Figure 7.37).
The case studies for crest deformations were sorted based on core width, core
material type and moisture content at placement. For embankments with very broad
earthfill cores the influence of the foundation was taken into consideration. For the
Chapter 7: Deformation behaviour of embankment dams Page 7.292

deformation versus time for the shoulders, the case studies have been sorted based on
the material type in the shoulder, and for rockfills its compaction rating.
The general trend of the deformation versus log time plots is:
• For settlement, near linear to slightly increasing rate with time (rate in terms of log
time) is the general trend. But, it is not unusual to have a slightly decreasing
settlement rate long-term for the crest. In some cases magnitudes of settlement are
large on first filling for the crest and upstream slope, particularly where the
upstream shoulder is susceptible to collapse settlement.
• For horizontal displacement a general trend is more difficult to define because of the
broader range in behaviour, but:
− A large percentage of the displacement generally occurs on first filling. For the
crest and downstream slope, this displacement is generally in a downstream
direction.
− Long-term, the trend of the displacement rate (rate per log time) approaches low
to near zero values, with fluctuations about the trend due to reservoir fluctuation.

Table 7.26: Figure references for post construction crest deformation


Core Core *2 Moisture Figure Reference
1 3
Width * Classification Content * Settlement Displacement
Dry placed Figure 7.53 Figure 7.68
CL/CH
Wet placed Figure 7.54 Figure 7.69
Thin to
Dry placed Figure 7.55 Figure 7.70
Medium SC/GC
Wet placed Figure 7.56 Figure 7.71
SM/GM Dry and wet placed Figure 7.57 Figure 7.72
CL/CH Mostly dry placed Figure 7.58 Figure 7.73
Thick
SC/GC/SM/GM Mostly dry placed Figure 7.59 Figure 7.74
Limited/negligible foundation influence Figure 7.60 Figure 7.75
Very Broad
Potentially significant foundation influence Figure 7.61 Figure 7.76
Note: *1 The terms used for core width classification are defined in Section 1.3.3.
*2 The core classification is to the Australian Soil Classification System.
*3 Dry placed defines cores where limited to negligible pore water pressures were developed
during construction and is generally applicable to placement more than 0.5% to 1% dry of
Standard optimum. Wet placed refers to cores where significant pore water pressures were
developed during construction. Refer to Section 7.4.1.4 for further discussion and definition
of the qualitative terms used.
Chapter 7: Deformation behaviour of embankment dams Page 7.293

Table 7.27: Figure references for post construction deformation of the embankment
shoulders
Shoulder Downstream Shoulder *2 Upstream Shoulder *2
Compaction
Material
Rating *1 Settlement Displacement Settlement Displacement
Type
Well compacted Figure F2.4 Figure F2.12 Figure F2.23 Figure F2.30
Reasonably to
Figure F2.5 Figure F2.13 Figure F2.24 Figure F2.31
Rockfill well compacted
Reasonably
Figure F2.6 Figure F2.14 Figure F2.25 Figure F2.32
compacted
Poorly compacted Figure F2.7 Figure F2.15 Figure F2.26 Figure F2.33
Gravels - Figure F2.8 Figure F2.16
Figure F2.27 Figure F2.34
Earthfills - Figure F2.9 Figure F2.17
No foundation
Very broad Figure F2.10 Figure F2.18 Figure F2.28 Figure F2.35
influence
core width
Foundation
(earthfill) Figure F2.11 Figure F2.19 Figure F2.29 Figure F2.36
influence
Note: *1 Refer to Section 1.3.3 for definitions of the compaction rating terms used.
*2 Figures are in Section 2 of Appendix F.

(b) Magnitude of post construction settlement


The magnitude of post construction settlements for the crest and shoulder regions are
presented at 3, 10 and 20 to 25 years after the end of embankment construction. Table
7.28 is a pointer to tables and figures within the chapter. Table 7.29 and Table 7.30
summarise the typical range of “normal” settlement magnitude for the crest and
shoulders respectively.
For the crest region, the data has been sorted based on core width, core material type
and placement moisture content, and indicates that:
• The post construction crest settlements are generally much smaller than the core
settlement during construction.
• Nearly all dams experience less than 1% crest settlement post construction for
periods up to 20 to 25 years and longer after construction.
• Most experience less than 0.5% in the first 3 years and less than 0.75% after 20 to
25 years.
• Smaller magnitude settlements are observed for dry placed clayey sands to clayey
gravels and dry to wet placed silty sands to silty gravels.
Chapter 7: Deformation behaviour of embankment dams Page 7.294

• A broader range of settlement magnitude is shown for clay cores, wet placed clayey
sand to clayey gravel cores, and embankments with very broad core widths.
• For zoned earth and rockfill dams, poor compaction of the rockfill is over-
represented for case studies at the larger end of the range of crest settlement.

The data for the shoulder regions indicates:


• Settlements in the order of 1 to 2% are observed for poorly compacted rockfills.
Greater settlements are observed for the dry placed, poorly compacted rockfills.
• For reasonably compacted rockfills the range of settlement is quite broad, from
0.1% up to 1.0%, but the number of cases is limited. Settlements toward the upper
range are observed for dry placed and/or weathered rockfills, where settlements due
to collapse compression are likely to be significant.
• Much lower settlements, generally less than 0.5 to 0.7% at ten years after
construction, are observed for well and reasonably to well compacted rockfills, and
compacted earthfills.
• Very low settlements (less than 0.25% at 10 years) are observed for embankments
with gravel shoulders.

(c) Long-term settlement rate


As previously discussed, post first filling the settlement rate (rate in terms of log time)
of SMPs on the crest and slopes is generally close to linear. For a number of case
studies the rate may increase slightly with time and for some case studies it may
decrease slightly with time. The long-term settlement rate for the case studies was
estimated assuming a linear relationship between settlement and log time, and over
periods of time representing “normal” reservoir operating conditions. For case studies
where the rate increased (or decreased) with time after first filling the estimate was
generally based on the later period of measurement. Table 7.31 is a pointer to tables
and figures related to long-term settlement rate within the chapter, and Table 7.29 and
Table 7.30 provide the typical range of long-term settlement rate for the crest and
shoulders respectively excluding possible outliers. The units of settlement rate are
percent settlement per log cycle of time (settlement as a percentage of the height from
the SMP to foundation level). The same units are used for the puddle core earthfill
embankments.
Chapter 7: Deformation behaviour of embankment dams Page 7.295

Table 7.28: References to post construction settlement magnitude tables and plots
Embankment Table Time After End of Construction (years) *1
Region Reference 3 years 10 years 20 to 25 years
Crest Table 7.19 Figure 7.46 Figure 7.47 Figure 7.48
Figure 7.49 Figure 7.50
Downstream shoulder Table 7.20 Figure F2.3
(Figure F2.1) (Figure F2.2)
Figure 7.51 Figure 7.52
Upstream shoulder Table 7.20 Figure F2.22
(Figure F2.20) (Figure F2.21)
Note: *1 Figures numbers with an ‘F’ prefix are in Section 2 of Appendix F.

Table 7.29: Embankment crest region, typical range of post construction settlement and
long-term settlement rate
Long-term Settlement
Core Properties Crest Settlement (%) *1, *2
Rate *1, *3
Core Moisture 20 to 25 Steady/Slow Fluctuating
Classn 3 years 10 years
Width content years Reservoir Reservoir
0.05 to 0.10 to 0.20 to
dry
Thin to 0.55 0.65 0.95
medium 0.04 to 0.08 to 0.20 to 0.04 to 0.50
CL/CH wet 0.09 to 0.57
0.75 0.95 1.10 (most < 0.26)
all (most 0.02 to 0.10 to
Thick 0.5 to 1.0
dry) 0.75 1.0
0.10 to 0.10 to
dry < 0.5
Thin to 0.25 0.40
medium 0.15 to 0.20 to
SC/GC wet < 1.1 0 to 0.26 0.06 to 0.37
0.80 1.10
all (most 0.05 to 0.10 to 0.10 to
Thick
dry) 0.20 0.35 0.45
Thin to 0.06 to 0.10 to < 0.5 to
SM/GM all < 0.10 0.03 to 0.21
thick 0.30 0.65 0.7
Very Broad Earthfill Cores - 0.0 to 0.0 to 0.05 to 0.07 to 0.70
0.08 & 0.44
most CL and dry placed 0.60 0.80 0.76 (most < 0.35)
Note: Classn = classification to Australian Soil Classification System
*1 excludes possible outliers.
*2 crest settlement as a percentage of the embankment height
*3 long-term settlement rate in units of % settlement per log cycle of time (settlement as a
percentage of dam height).
Chapter 7: Deformation behaviour of embankment dams Page 7.296

Table 7.30: Embankment shoulder regions, typical range of post construction settlement
and long-term settlement rate
Downstream Shoulder *1 Upstream Shoulder *1
Material Compaction
Settlement (%) *2 Settlement Settlement *2 Settlement
Type Rating 3
3 years 10 years Rate * 3 years 10 years Rate *3
well 0.05 to 0.05 to 0.0 to 0.33 0.10 to 0.10 to 0.05 to 0.70
0.35 0.55 0.60 0.70 (most < 0.50)
reasonably to < 0.30 < 0.50 0.04 to 0.31 0 to 0.55 0.10 to 0.10 to 0.56
well (most > 0.15) 0.60 (most < 0.50)

Rockfill reasonable 0.20 to 0.10 to 0.10 to 1.0 < 0.70 - *5 < 0.55
1.0 1.0
poor 0.10 to ? 0.15 to ? 0.10 to 0.25 0.10 to 0.15 to 0.10 to 0.82
*5 *5 1.05 1.20 (most < 0.60)
poor – dry *4 0.15 to 0.30 to 0.20 to 0.75 0.15 to 0.20 to
1.60 2.00 1.35 1.6
Gravels - < 0.15 < 0.25 0.02 to 0.065 < 0.15 < 0.25 < 0.21

Earthfills - 0.0 to 0.0 to 0.0 to 0.40 0.05 to 0.10 to 0.10 to 0.60


0.40 0.70 0.60 0.70
Note: *1 Excludes possible outliers.
*2 Settlements quoted are a percentage of the height from the SMP to foundation level.
*3 The long-term settlement rates are in units of % settlement per log cycle of time (settlement
as a percentage of the height from the SMP to foundation level).
*4 For the dry placed and poorly compacted rockfills, a large range in settlements is observed.
For rockfills placed in dry climatic regions settlements are likely to be toward the upper end
of the range.
*5 insufficient data.

Table 7.31: References to tables and figures of long-term settlement rate


Embankment Table Figure
Comment
Region Reference Reference
Crest Table 7.21 Figure 7.62 Figure 7.62 is for zoned embankments
and with thin to thick core widths, and Figure
7.63 is for embankments with very broad
Figure 7.63
core widths.
Downstream shoulder Table 7.22 Figure 7.78
Upstream shoulder Table 7.23 Figure 7.83

Reservoir fluctuation was found to influence the long-term settlement rate for the crest,
particularly for the zoned embankments of thin to thick core width with permeable
upstream shoulder fill (i.e. rockfill or gravels). Generally, greater long-term settlement
rates were measured for case studies with fluctuating reservoir levels. The influence of
Chapter 7: Deformation behaviour of embankment dams Page 7.297

reservoir fluctuation on the upstream shoulder could not be clearly identified from the
available data and it had limited to negligible influence on the downstream shoulder.
Two qualitative categories were used for reservoir operation, “fluctuating” for
reservoirs subject to a seasonal (usually annual) or regular (more than once per year)
drawdown typically greater than about 10 to 15% of the maximum height of the
embankment, and “steady” or “slow”.
Apart from the influence of reservoir operation, the data for the crest indicated:
• For zoned embankments with thin to thick core widths:
− Core material type has an influence on the long-term crest settlement rate. In
general, earthfill cores of silty sands to silty gravels tend to have low long-term
settlement rates and clay earthfill cores, on average, tend to have high long-term
crest settlement rates. Clayey sand to clayey gravel cores show intermediate
rates.
− The core width and compaction moisture content appear to have little to no
recognisable influence on the long-term crest settlement rate. Their influence is
probably over-shadowed by other factors.
− Embankments with rockfills susceptible to large deformations due to collapse
compression have high long-term settlement rates, generally greater than 0.4%
per log cycle of time, indicating these embankments are more susceptible to
softening and long-term degradation.
• For embankments with very broad width earthfill, zones the long-term settlement
rate shows a broad variation in range. Reservoir fluctuation does influence the long-
term settlement rate for those embankments with more permeable earthfills.

The data for the embankment shoulders indicates:


• The long-term settlement rate of both the up and downstream shoulder is generally
less than about 0.4% per log cycle of time for most cases.
• Higher rates are observed for zoned earth and rockfill dams with rockfills
susceptible to large deformations due to collapse compression, particularly in the
downstream shoulder.
• Very low rates for zoned embankments with gravel shoulders.
Chapter 7: Deformation behaviour of embankment dams Page 7.298

(d) Post construction horizontal displacement


As previously indicated, a large portion of the horizontal displacement generally occurs
on first filling, particularly for zoned earthfill and earth-rockfill dams with permeable
fills in the upstream shoulder. Additional data (other than the displacement versus time
plots) is presented for horizontal displacement, including:
• Lateral displacement on first filling:
− Table 7.17 and Figure 7.38 for displacement of the crest region;
− Figure 7.39 for the downstream shoulder region; and
− Figure 7.40 for the upstream shoulder region
• Lateral displacement of the crest post first filling (Figure 7.77).

The case study records on crest displacement during first filling show that the
typical range for most case studies is from 50 mm upstream to 300 mm downstream. A
number of trends were evident sorting the data based on core width, core material type
and the material type and compaction rating of the downstream shoulder. They were:
• For most groups, the displacement on first filling was downstream and less than 0.1
to 0.2% of the embankment height (less than 100 to 200 mm).
• Greater displacements (from 0.2% to 0.6% and up to almost 1% of the embankment
height) were observed for central core earth and rockfill dams of thin to medium
core width with:
− Rockfills susceptible to large deformations due to collapse compression in the
downstream shoulder.
− Silty sand to silty gravel earthfill cores.

These findings suggest that large deformations in the downstream shoulder, possible
due to collapse compression on wetting from rainfall, significantly influence the crest
displacement for embankments of thin to medium core width. This was confirmed by
the observation that those embankments with large crest displacements on first filling
also experienced large downstream displacements of the downstream shoulder. The
correlation to silty sand and silty gravel cores is possibly indicative of greater
magnitude increases of lateral stress in the downstream shoulder on first filling for these
embankment types.
Chapter 7: Deformation behaviour of embankment dams Page 7.299

The displacement of the downstream shoulder on first filling had many similarities
to the crest behaviour on first filling:
• Displacements of the downstream shoulder were less than 0.1 to 0.2% of the
embankment height for embankments with very broad earthfill cores, zoned
embankments with compacted gravels or earthfills in the downstream shoulder, and
zoned embankments with wetted (and generally compacted) rockfills in the
downstream shoulder.
• Greater magnitude displacements were generally observed for zoned earth and
rockfill dams with dry placed rockfills in the downstream shoulder:
− Up to 0.25 to 0.30% of the dam height for dry placed and well and reasonably to
well compacted rockfills
− More than 0.2 to 0.25% and up to 0.85% of the dam height for dry placed and
poor and reasonably compacted rockfills.

For the upstream shoulder region, displacements on first filling ranged from 200 to
300 mm upstream to 300 mm downstream.
Crest displacements for the period from post first filling to 10’s of years after
construction are generally of smaller magnitude than the displacement on first filling,
particularly for zoned earth and earth-rockfill embankments with thin to medium core
widths. Regardless of the embankment type though, the general range of displacement
post first filling (for at least 5 years post first filling and up to 40 to 50 years) is quite
small, ranging from 35 mm upstream to 100 to 150 mm downstream.
The displacement of the downstream slope post first filling is more erratic than for
the crest, even though for a large number of the case studies the displacement post first
filling displacement is of small magnitude. In terms of the total magnitude of
displacement in a downstream direction since end of construction:
• For embankments with very broad core widths, total displacements generally range
from 0.05 to 0.30% of the embankment height at 25 to 45 years after end of
construction.
• For zoned earthfill embankments, total displacements are up to 0.15 to 0.20% of the
embankment height at 20 to 30 years after end of construction.
• For central core earth and rockfill dams:
− For well and reasonably to well compacted rockfills, displacements of up to
0.20% are measured at 10 to 20 years after end of construction. Where the
Chapter 7: Deformation behaviour of embankment dams Page 7.300

rockfill has been dry placed or is of relatively high compressibility,


displacements can be greater; up to 0.25 to 0.40 % of the embankment height.
− For reasonably and poor compacted rockfills, displacements up to 1.0 to 1.6% of
the embankment height can occur long-term, particularly for dry placed rockfills
susceptible to large deformations due to collapse compression.

7.14.4 PUDDLE CORE EARTHFILL DAMS

The predictive methods for puddle core earthfill dams are appropriate to the long-term
deformation behaviour of the embankment, many tens of years after construction. The
following presents a summary of the analysis and discussion from Section 7.7.3, of
which Section 7.7.3.6 provides a useful summary of the factors affecting the long-term
deformation behaviour of puddle core earthfill dams.
The records show that the post construction crest settlement of puddle core earthfill
embankments is significant, ranging from 1% up to 8% to 14% of the dam height after
more than 100 years. The older (pre 1900) embankments generally show the greater
magnitude post construction crest settlement and the more recent embankments
(constructed in the 1930’s to 1950’s) generally show crest settlements of lesser
magnitude. A large proportion of the settlement in the older embankments occurred
during and shortly after the period of first filling due mainly to collapse compression on
wetting of the poorly compacted earthfill supporting the puddle core and yielding on
drawdown.
The long-term settlement rate (SLT) of the embankment crest (under normal reservoir
operating conditions) was significantly affected by the magnitude of reservoir
fluctuation and the permeability of the earthfill zone upstream of the puddle core. Table
7.32 provides guidelines for estimation of the long-term crest settlement under normal
reservoir operating conditions. The influence of reservoir fluctuation and earthfill
permeability over-shadowed other factors that probably influence the long-term crest
settlement including dam height, age and earthfill material type. For several dams the
long-term crest settlement rate was intermediate to the ranges given. For Hope Valley
dam, part of the reason was thought to be the variable pore water pressure response
(ranging from 25 to 100%) in piezometers in the earthfill zone upstream of the puddle
core to fluctuations in reservoir level.
Chapter 7: Deformation behaviour of embankment dams Page 7.301

Table 7.32: Predictive methods of long-term crest settlement and displacement under
normal reservoir operating conditions for puddle core earthfill embankments
Reservoir Response to No. Crest Settlement Crest
Operation Drawdown of the Cases Prediction Displacement
*1 Earthfill Zone *3
Upstream of the
Puddle Core *2
Steady (“steady state”)
negligible to minor 7 SLT = 0.4 to 1.0 % *4 negligible
(or “steady state”)
SLT = 4.5 to 7.4 %

Fluctuating
near full response Figure 7.98 for estimation 3 to 5 mm/year
4
(i.e. permeable earthfill) of permanent settlement (downstream)
based on the magnitude
of the drawdown
Notes: *1 Fluctuating defined as reservoir subject to annual drawdowns generally greater than 10 to
20% of the embankment height
*2 Refer Figure 7.93 for definitions of “steady state” conditions.
*3 crest displacement records only available for very few case studies.
*4 SLT is the long-term crest settlement rate in units of settlement as percentage of embankment
height per log cycle of time. Individual values for the case studies are given in Table 7.24.

The methods based on historical performance of similar dams are very approximate
because of the limited number of case studies from which they have been derived and
the assumptions made regarding the reservoir operation and permeability of earthfill
zones upstream of the core. The methods should only be as a general guide in
consideration of these factors.
For several embankments, large permanent crest settlements (0.20 to 0.52% of the
embankment height) were measured during abnormally or historically large drawdown
events (Tedd et al 1997b). In terms of prediction of deformation during these abnormal
events, the only available method would seem to be fully coupled finite element
modelling (FEM) due to the complexity of the processes and interaction between the
different elements within the embankment. But, this is not without difficulties due to
the lack of known information on material properties as well as the need to consider
stress history.
Available records on the long-term deformation of the shoulders and displacement
of the crest are limited. Some preliminary guidance on possible magnitudes of long-
term deformation under normal reservoir operating conditions is:
• For displacement of the crest and downstream slope:
Chapter 7: Deformation behaviour of embankment dams Page 7.302

− Under “steady state” conditions long-term rates of displacement are negligible.


− For fluctuating reservoir conditions and permeable upstream earthfill zones,
long-term rates of displacement of up to 3 to 5 mm/years in a downstream
direction have been recorded.
• The long-term settlement rate for the shoulders is:
− Of similar magnitude to the crest for “steady state” conditions (refer Table 7.32).
− Of lesser magnitude than the crest for fluctuating reservoir conditions and
permeable upstream earthfill zones.

7.15 CONCLUSIONS
The main objective of this chapter was to develop methods for identification of
potentially “abnormal” deformation behaviour of embankment dams from field
monitoring records. The methods were predominantly developed from initially defining
what is “normal” deformation behaviour for a particular embankment type in
consideration of the imposed stress conditions, and the strength properties and stress-
strain relationship of the materials. Potentially “abnormal” deformation behaviour was
then broadly identified where the rate, magnitude or trend of the deformation behaviour
differed from that of the “norm”. Implicit in the methods are concepts such as
consolidation, creep under stress conditions and collapse compression on wetting. Once
the deformation behaviour for a case study was identified as potentially “abnormal”,
further analysis was undertaken (where records were available) for evaluation of the
possible mechanism/s causing the observed deformation behaviour. Sections 7.8 to 7.12
of the chapter deal with the identification and evaluation of “abnormal” deformation
behaviour from the case study data, Section 7.13 providing a summary.
The database comprises some 134 embankments and included the following
embankment types:
• Zoned earth and rockfill embankments, most of which were central core earth and
rockfill embankments.
• Zoned earthfill embankments
• Rolled earthfill embankments, including homogeneous earthfill, earthfill with filters
and earthfill with rock toe.
• Puddle core earthfill embankments.
Chapter 7: Deformation behaviour of embankment dams Page 7.303

The deformation monitoring records analysed included:


• During construction – mainly internal vertical deformation of the core, but also the
lateral core deformation of zoned earth and rockfill dams with thin to medium core
widths.
• Post construction – SMPs on the crest and slopes of the embankment, and the
internal vertical deformation of the earthfill core.

From such a broad database of embankment dams it has been possible to develop
methods for prediction and evaluation of the deformation behaviour during and post
construction from the main types of deformation monitoring records analysed. Section
7.14 provides a summary of these methods with reference to pertinent tables and figures
within the chapter.
The methods are considered to be an improvement on currently available methods
for the embankment types considered. The influence of material type and placement
methods, embankment zoning geometry, embankment height, reservoir operation
amongst other factors have been considered.
Chapter 8: Conclusions and Recommendations Page i

TABLE OF CONTENTS
8.0 CONCLUSIONS AND RECOMMENDATIONS...................... 8.1
8.1 CONCLUSIONS ................................................................................................. 8.1
8.1.1 Landslides in Soil Slopes ...........................................................................8.1
8.1.2 Deformation Behaviour of Embankment Dams......................................... 8.4
8.2 RECOMMENDATIONS FOR FURTHER R ESEARCH ............................................... 8.6
Chapter 8: Conclusions and Recommendations Page 8.1

8.0 CONCLUSIONS AND RECOMMENDATIONS

8.1 CONCLUSIONS
This thesis is a study of the pre and post failure deformation behaviour of landslides in
cut, fill and natural soil slopes, and of the deformation behaviour of embankment dams.
The primary objective of the research has been to further develop the understanding of
the deformation behaviour of landslides and embankment dams, and then to improve
predictive methods for use in quantitative risk assessment of landslides, and the
evaluation and prediction of the deformation of embankment dams.
The research has been undertaken within the general framework of the geotechnical
characterisation system of slope movements (as described by Leroueil et al. 1996),
mainly concentrating on deformation related aspects.

8.1.1 Landslides in Soil Slopes

From the study of the deformation behaviour of landslides in soil slopes, it has been
possible to identify the factors contributing to the pre and post failure behaviour of the
landslide groups studied.
The progression of the initial slope failure to a “rapid” debris flow or “slow” intact
slide post failure (for example) is primarily dependent on the mechanics of the failure,
slope geometry, geometry of the surface of rupture and material properties of the slide
mass. It is not usually controlled by the trigger mechanism/s for the processes leading
up to slope instability.

(a) Pre-failure deformation behaviour


The mechanics of failure are significant in controlling the pre-failure deformation
behaviour of a landslide. A progressive failure mechanism is present in a large number
of slope failures in soil slopes. For soils that are highly strain weakening on shearing,
progressive failure may develop very rapidly from a localised failure condition in the
slope (e.g. in the case of flow liquefaction) and show little observable pre-failure
deformation. In other cases the progressive strain weakening in the slope may develop
much more slowly, and for these cases appropriately located and monitored deformation
Chapter 8: Conclusions and Recommendations Page 8.2

instrumentation would, with proper interpretation, allow identification of the impending


failure.
Other factors also influence the pre-failure deformation behaviour, and these vary
from one slide group to the next. For example, the permeability (and therefore particle
size) is thought to affect the pre-failure deformation behaviour of slides in contractive
soils. For other slide groups, the requirement for internal shearing to form a
kinematically admissible slide mechanism (e.g. in defect controlled compound slides in
completely weathered rock masses or stiff jointed clays) and brittleness on the slide
margins can affect pre-failure deformation. For other cases again, it may not be
worthwhile attempting to monitor pre-failure deformation because the trigger of failure
is related to climatic conditions, and the slope in question may undergo a reduction in
stability from adequate to limiting over a period of hours.
Appropriately located and interpreted deformation monitoring instrumentation
would (in most cases) identify the impending failure for slides in embankments of soft
ground, larger cut slopes in heavily over-consolidated high plasticity and other clays,
embankment dams, flow slides in granular waste spoil piles, and defect controlled slides
in weathered rock masses, but for these, mainly for compound slides.

(b) Post failure deformation behaviour


Post failure, the potential for strain weakening strongly influences the deformation
behaviour of the failed slide mass. The concept of conservation of energy, and the
consideration of the potential for strain weakening of the slide mass within this
framework, is a useful method for qualitative assessment of the post failure deformation
behaviour. For landslides in soils that are distinctively strain weakening on shearing on
the surface of rupture or have a high degree of internal brittleness associated with the
slide release mechanism (e.g. toe buttressing), a significant amount of the potential
energy of the slide can be available for kinetic energy and dis-aggregation of the slide
mass. This would lead to acceleration of the slide mass to “rapid” post failure velocities
resulting in potentially large travel distances. At the other end of the scale, slope
failures that are triggered by applied load, such as in the case of embankment
construction, and for which the soils are not significantly strain weakening on shearing,
post failure velocities will generally be slow, the travel distance of limited extent and
the slide mass will remain virtually intact during travel.
Chapter 8: Conclusions and Recommendations Page 8.3

From the case study analysis, the factors that strongly influenced the post failure
deformation behaviour of the slide mass were; the mechanics of failure, the source area
slope angle, the down-slope geometry, the slope failure geometry (including the
orientation of the surface of rupture) and the stress-strain properties of the soil type.
“Rapid” post failure velocity of the slide mass is generally developed where one or
more of the following is present:
• Failure was within saturated granular soils that are contractive on shearing and
within which a flow liquefaction condition occurred;
• A high degree of brittleness is associated with the slide mechanics or release on the
margins;
• Steep source area slope (for all soil types); and
• Steep slopes immediately downslope of the initial slide.

For these slides, often the slide mass disaggregates and develops into a debris flow.
Post failure travel distances are generally large and at velocities in the order of metres to
10’s of metres per second. For slides in steep natural slopes in dilative soils, the soil
type was a significant factor in the ability of the slide mass to transform into a “rapid”
debris flow, with low plasticity or non-cohesive silty to sandy and gravelly soils being
the most susceptible.
“Slow” slides that generally remain intact on sliding are generally characterised by a
relatively flat basal slide plane or a rotational surface of rupture that passes through the
horizontal. Post failure travel distances tend to be of limited extent and at “slow” post
failure velocities.
Methods for assessment of the post failure travel distance have been developed for
landslides of both “rapid” and “slow” post failure velocity.
For landslides of “rapid” post failure velocity it is apparent that the use of slide
volume alone does not allow reliable predictions of the travel distance angle (and travel
distance). Improvements in predictions for a number of slide groups have been
developed from consideration of a combination of material type, mechanics of failure,
slope geometry, travel path confinement and/or slide volume. An appropriate degree of
caution should be used in application of the methods due to the uncertainty in the
methods, and they should be applied to conditions similar in which they have been
derived. In all cases, it is best to calibrate the models with case studies for the area
under study so that local geological conditions and climatic factors can be allowed for.
Chapter 8: Conclusions and Recommendations Page 8.4

For landslides that generally remain intact and are of “slow” post failure velocity, it
is demonstrated that the Khalili et al (1996) methods provides a reasonable means for
estimation of the post failure deformation of the slide mass. The methods have been
developed for rotational slides, but can be applied to slides of compound geometry.
They may not be suitable for analysis of slides where internal brittleness in the slide
mechanics is significant.

8.1.2 Deformation Behaviour of Embankment Dams

(a) Methods for Prediction of “Normal” Deformation Behaviour


Improved methods for prediction and/or evaluation of the deformation behaviour of an
embankment dam (during and post construction) have been developed from an
extensive database of case studies. The methods take into consideration the influence of
material type and placement methods, embankment zoning geometry, embankment
height, and reservoir operation, amongst other factors.
From the case study data during embankment construction, methods for predicting
the modulus of compacted rockfill have been developed that take into account the intact
rock strength and the particle size distribution of the rockfill, correcting the vertical
stress for embankment shape and cross valley influence. Guidelines have also been
developed for estimation of the stress-strain relationship of compacted rockfills for use
deformation prediction during construction.
The deformation behaviour during construction of compacted earthfills is dependent
on the soil type, its moisture content at placement (relative to Standard optimum
moisture content), the embankment type, and the material type and compressibility
properties of the supporting shoulders. For “dry” placed earthfill cores (placed drier
than about 0.5% dry of Standard optimum) of thick to very broad width, it is
demonstrated that the assumption of one-dimensional compression and use of elastic
solutions are reasonable for prediction of the deformation behaviour under the
embankment centreline. Guidelines on the apparent secant moduli of these compacted
earthfill types from the case study data have been developed.
For “wet” placed earthfill cores (placed wetter than about 0.5% dry of Standard
optimum) of thin to medium width, a significant proportion of the deformation during
construction is due to undrained plastic type deformations. For these embankments, the
compressibility properties of the supporting shoulders significantly influence the
Chapter 8: Conclusions and Recommendations Page 8.5

deformation of the core, with strains of large magnitude observed for shoulders of high
compressibility (e.g. poorly compacted rockfills or compacted weathered rockfills).
Post construction, methods have been developed for evaluation / prediction of the
crest settlement and face slab deformation of concrete face rockfill dams. The method
for crest settlement is by summation of the settlement due to first filling and the time
dependent deformation components. The time dependent component for compacted
rockfills takes into consideration material type (i.e. rockfills from quarried rock or use
of gravels), the intact strength of rock (for quarried rockfills) and embankment height.
Methods for evaluation / prediction of the post construction deformation of
embankment dams (both settlement and lateral displacement) have been developed for
the crest and shoulder regions. The data is presented in several formats; as deformation
versus log time, as snap shots in time at 3, 10 and 20 to 25 years after the end of
construction (end of first filling for horizontal displacement), and long-term settlement
rate. The factors that influence the deformation behaviour vary for each region of the
embankment, and also vary for the component of deformation (i.e. settlement or
displacement). The factors considered in the methods include: the material type,
placement moisture condition and width of the main earthfill zone; the material type and
placement method of the shoulders; the embankment height; the reservoir operation and
the foundation influence. Central core earth and rockfill embankments with rockfill
shoulders that are susceptible to large deformations due to collapse compression are
often over-represented at the high end of the rate or magnitude of deformation.
For all case studies analysed, zero time for the analysis of the post failure
deformation has been established at the end of construction. For concrete face rockfill
dams this was defined as the end of main rockfill construction.

(b) Methods for Identification of “Abnormal” Deformation Behaviour


Existing general methods or guidelines for identification of “abnormal” deformation
behaviour of embankment dams are limited. In general, practitioners rely on their
experience, limit equilibrium analysis and simplified numerical methods.
Improvements in this area have been made from the case study analysis and
consideration of the mechanics controlling the “abnormal” deformation behaviour.
Guidelines have been developed to assist in the identification of potentially “abnormal”
deformation behaviour, both during and post construction. The methods of assessment
are based on the identification of outliers to the general trend of deformation behaviour
Chapter 8: Conclusions and Recommendations Page 8.6

for a given embankment type and consideration of such factors as the core material
type, its moisture content at placement, its width, the type and placement method of the
supporting shoulders (for zoned embankments) and reservoir operation.
It has been demonstrated that many cases of larger than “normal” deformation
behaviour are related to lack of support of the dam core from poorly compacted or
dumped rockfill shoulders that, in general, are susceptible to large deformations due to
collapse compression on wetting. It has been shown for some of these dams that, over
time, spreading, cracking and softening of the core has progressively developed. In
some these cases (as well as for other case studies), shear surfaces have developed in the
core despite the overall factor of safety being adequate. Large or historic reservoir
drawdowns have been shown to trigger further shear movement along the existing shear
surface in the core in a number of case studies.
There are however case studies where “abnormal” deformation behaviour is related
to marginal stability of the whole or parts of the embankment, as is the case for
Eppalock, Djatiluhur and the closure section at Belle Fourche dams. Guidelines for
identification of a marginal stability condition have been developed. The considerations
include: evidence of shear development; “abnormally” large incremental deformations
as layers are placed (during construction); the persistence of longitudinal and diagonal
cracking in the core; continued acceleration of the deformation on successive large
drawdowns of the reservoir; and the tertiary creep deformations under constant stress
conditions.
A critical component of any deformation monitoring program is the foresight to
envisage the potential modes of marginal stability, install appropriate monitoring
equipment to detect the condition and take measurements within a timing framework
that encapsulates the envisaged period of likely minimal stability (i.e. the upstream
slope during drawdown or the downstream slope during a reservoir full condition). The
deformation behaviour can then be linked to analysis of stability using limit equilibrium
and/or finite element methods.

8.2 RECOMMENDATIONS FOR F URTHER RESEARCH


The overall analysis of the pre and post failure deformation of landslides in soil slopes,
and of the deformation behaviour of embankment dams in this thesis is case study
driven. What has been achieved here has been limited by the quantity and quality of the
Chapter 8: Conclusions and Recommendations Page 8.7

case study data despite an extensive search in Australia, Europe, USA/Canada and Hong
Kong. There is a particular lack of pre failure deformation data for landslides in cuts,
fills and natural slopes. Only for embankment dams is there extensive data, but most of
this is not related to the deformation behaviour leading up to slope instability. Post
failure data is often of poor quality, although for several of the slide types studied good
quality data was available for a number of case studies, particularly the data from Hong
Kong.
In the area of post failure deformation behaviour of landslides, it is considered that
the slide groups that would most readily benefit from further case study analysis
(mainly because of the prevalence of case study records available in the international
community) is “rapid” landslides from failures in steep natural slopes and steep cut
slopes. “Rapid” debris flows from failures in steep natural slopes occur in regions all
over the world and are widely reported in the literature. They were not analysed here
because only limited details on the individual case studies were reported in the
published literature.
Other areas from the study of the deformation behaviour of landslides in soil slopes
for which further research would be beneficial include:
• Further verification of the post failure analysis of intact landslides using the Khalili
et al (1996) model and/or other lumped mass models. The guidelines established are
based on a limited number of case studies.
• Dynamic analysis of the post failure deformation behaviour of “rapid” landslides.
This has been done for several classes of slope including flow slides in waste spoil
piles in British Columbia (Golder Associates 1995) and “rapid” landslides in Hong
Kong (Hungr Geotechnical Research 1998; Ayotte and Hungr 1998). This could be
extended to “rapid” landslides in other areas and the profession would benefit
greatly from having a reliably calibrated commercially available program to model
flows.
• Pre-failure deformation of cuts and fills to develop further the guidelines for
identification of an impending failure condition would be valuable.

For embankment dams, the most significant area of further research is considered to
be in the systematic numerical modelling of individual case studies using constitutive
models that realistically describe the stress-strain behaviour of earthfills and rockfills.
Some of the concepts raised in this thesis rely on simplified modelling using a Mohr-
Chapter 8: Conclusions and Recommendations Page 8.8

Coulomb linear-elastic perfectly plastic model. Verification of these concepts by more


rigorous numerical modelling would be of great benefit. This should include the use of
coupled models, and constitutive models that consider the effects of partial saturation on
the stress-strain relationship of fine-grained earthfills, as well as the non-linearity of the
strength and compressibility of earth and rockfills, and stress path dependency.
References Page i

TABLE OF CONTENTS

9.0 REFERENCES ............................................................................ R1


References Page R1

9.0 REFERENCES
Aas, G. (1981) Stability of natural slopes in quick clays. Proceedings of the Tenth
International Conference on Soil Mechanics and Foundation Engineering, San
Francisco, Balkema, Rotterdam, Vol. 3, pp. 333-338.
Aas, G., Lacasse, S., Lunne, T. and Höeg, K. (1986) Use of in situ tests for foundation
design in clay. Proceedings of the ASCE Conference on Use of In Situ Tests in
Geotechnical Engineering, (Clemence ed.) Blacksburg, Virginia, ASCE, New
York. pp. 1-30.
Abadjiev, C.B. (1994) Safety assessment and stability improvements of the upstream
slope of earth dams. Proceedings of the 18th International Congress on Large
Dams, Durban, ICOLD. pp. 261-273 (Q.68 R.20).
Abele, G. (1974) Bergsturze in den Alpen (in German). Wissenschaftliche
Alpenvereinshefte, No. 25, Munich.
Alberro, J. (1972) Stress-strain analysis of El Infiernillo dam. Proceedings, ASCE
Speciality Conference on Performance of Earth and Earth Retaining Structures,
Purdue University, Indianna, ASCE, New York. Vol. 1 (Part I) pp. 837-852.
Alberro, J.A. and Moreno, E. (1982) Interaction phenomena in the Chicoasen dam:
construction and first filling. Proceedings of the 14th International Congress on
Large Dams, Rio de Janiero, ICOLD. pp. 183-202 (Q.52 R.10).
Almeida, M.S.S. (1982) The undrained behaviour of the Rio de Janiero clay in the light
of critical state theories. Report No. TR 119 ISSN 0309 - 7439, Cambridge
University Engineering Department (CUED/D-Soils).
Almeida, M.S.S. and Ramalho-Ortigao, J.A. (1982) Performance of finite element
analyses of a trial embankment on soft clay. Proceedings, International Symposium
on Numerical Models in Geomechanics, (Dungar, Pande & Studer ed.) Zurich,
Balkema, Rotterdam. pp. 548-556.
Alonso, E., Fry, J.J. and Isambert, F. (1997) Rupture et confortment du barrage en argile
humide de Mondely et retour d'experience. Nineteenth International Congress on
Large Dams, Florence, ICOLD. Vol. 4, pp. 711-733 (Q75 R49).
Alonso, E.E. and Oldecop, L.A. (2000) Fundamentals of rockfill collapse. Proceedings
of the Asian Conference on Unsaturated Soils (Unsat-Asia 2000), (Rahardjo, Toll
and Leong ed.) Singapore, Balkema, Rotterdam. pp. 3-13.
References Page R2

Alvarez, A. and Bravo, G. (1976) A composed-core rockfill dam: the Canales dam.
Proceedings of the 12th International Congress on Large Dams, Mexico City,
ICOLD. Vol. 4, pp. 1077-1097 (C.20).
Amaya, F. and Marulanda, A. (1985) Golillas dam - design, construction and
performance. Proceedings of the Symposium on Concrete Face Rockfill Dams -
Design, Construction and Performance, (Cooke and Sherard ed.) Detroit,
Michigan, ASCE New York. pp. 98-120.
Amaya, F. and Marulanda, A. (2000) Columbian experience in the design and
construction of concrete face rockfill dams. In Concrete Face Rockfill Dams, J.
Barry Cooke Volume (Mori, Sobrinho, Dijkstra, Guocheng and Borgatti ed.),
Beijing, pp. 89-115.
Andong Dam Construction Office (1976) Development of new core material in zone fill
dam, use of decomposed granite in Andong dam. Proceedings of the 12th
International Congress on Large Dams, Mexico, ICOLD. pp. 731-748 (Q.44 R.38).
Andrus, R.D. and Stokoe, K.H. (1997) Liquefaction resistance based on shear wave
velocity. Proceedings of the NCEER Workshop on Evaluation of Liquefaction
Resistance of Soils, (Youd and Idriss ed.) Salt Lake City, Utah, National Centre for
Earthquake Engineering Research. pp. 89-128.
Anon (1918) Reconstruction of a retaining wall on the Great Central Railway. The
Engineer, Vol. 126, pp. 536-537.
Aphaiphuminart, S., Chanpayom, O., Mahasandana, T., Bhucharoen, V. and Pinrode, J.
(1988) Design, construction and performance - Khao Laem dam. Proceedings of
the 16th International Congress on Large Dams, San Francisco, ICOLD. pp. 95-
114 (Q.61 R.6).
Aulitzky, H. (1989) The debris flows in Austria. Bulletin of the International
Association of Engineering Geology, Bulletin No. 40, pp. 5-13.
Australian Capital Territory Electricity and Water (1993) ACT dam safety: Googong
dam surveillance report. Report No. DSU 93/04, Dam Safety Unit.
Australian Capital Territory Electricity and Water (1994) ACT dam safety: Corin dam
surveillance report. Report No. DSU 94/08, Dam Safety Unit.
Ayotte, D. and Hungr, O. (1998) Runout analysis of debris flows and debris avalanches
in Hong Kong, Final Report. Geotechnical Engineering Office, Hong Kong.
Ayoubian, A. and Robertson, P.K. (1998) Void ratio redistribution in undrained triaxial
extension tests on Ottawa sand. Canadian Geotechnical Journal, Vol 35, pp. 351-
359.
References Page R3

Bacon, G.A. (1969) Observed behaviour of dam embankments instrumented by the


Snowy Mountains Authority. ANCOLD Bulletin, Vol. 29 (October), pp. 40-54.
Bacon, G.A. (1999) Personal communication with Mr. G. Bacon on the construction of
Blowering dam.
Baker, R. (1972) Report of inspections and tests for Scotts Peak dam. Laboratory
Report 4164-1, Hydro-Electric Commission Tasmania, Civil Engineering
Laboratories.
Ballard, R., Smith, B. and Von Thun, L. (1981) Back analysis of the upstream slide
near Station 135 at San Luis dam (unpublished report). Interim Report No. 1,
United States Bureau of Reclamation, Division of Design, Embankment Dams
Section.
Banks, D., Strohm, W.E., De Angulo, M. and Lutton, R.J. (1975) Study of clay shale
along the Panama Canal. Report 3: Engineering analysis of slides and strength
properties of clay shales along the Gaillard Cut. Report 5-70-9, U.S. Army Eng.
Waterways Experiment Station, Vicksburg.
Banks, D.C. and Strohm, W.E. (1974) Calculations of rock slide velocities.
Proceedings, Congress ISRM, Denver, Vol. 2, pp. 839-847.
Banks, J.A. (1948) Construction of Muirhead reservoir, Scotland. Proceedings of the
2nd International Conference on Soil Mechanics and Foundation Engineering,
Rotterdam, Vol. 2, pp. 24-31.
Bar-Shany, M., Korlath, G. and Zeitlae, J.G. (1957) The use of fat clay in dam
construction in Israel. Proceedings of the 4th International Conference on Soil
Mechanics and Foundation Engineering, London, Butterworths Scientific
Publications, London. Vol. 2, pp. 273-277.
Baumann, P. (1939) Discussion on Galloway paper: The design of rock-fill dams.
Transactions, ASCE, Vol. 104, pp. 39-40.
Baumann, P. (1958) Rockfill dams: Cogswell and San Gabriel dams. A.S.C.E., Journal
of the Power Division, Vol. 84 (PO3), pp. 1687-1 to 33.
Baumann, P. (1964) Limit height criteria for loose-dumped rockfill dams. Proceedings
of the 8th International Congress on Large Dams, Edinburgh, ICOLD. Vol. 3, pp.
781- (Q.31 R.13).
Baziar, M.H. and Dobry, R. (1995) Residual strength and large deformation potential of
loose silty sands. A.S.C.E., Journal of Geotechnical Engineering, Vol 121 (12), pp.
896-906.
References Page R4

Becker, D., Crooks, J.H.A., Jeffries, M.G. and McKenzie, K. (1984) Yield behaviour
and consolidation, II: Strength gain. Proceedings of the ASCE Symposium on
Sedimentation Consolidation Models, Prediction and Validation, (Yong &
Townsend ed.) San Francisco, ASCE, New York. pp. 382-398.
Been, K. and Jefferies, M.G. (1985) A state parameter for sands. Geotechnique, Vol 35
(2), pp. 99-112.
Been, K., Jefferies, M.G. and Hachey, J. (1991) The critical state of sands.
Geotechnique, Vol 41, (3), pp. 365-381.
Beene, R.R.W. (1967) Waco dam slide. Journal of the Soil Mechanics and Foundations
Division, ASCE, Vol. 93 (SM4), pp. 35-44.
Benassini, A., Casales, V., Hungsberg, U., Canales, R. and Esquivel, R. (1976) Mexican
National Committee on Large Dams; General paper. Proceedings of the 12th
International Congress on Large Dams, Mexico, ICOLD. Vol. 4, pp. 609-660
(G.P.9).
Bentley, S.P. and Smalley, I.J. (1984) Landslips in sensitive clays. In Slope Instability
(Brunsden and Prior ed.), John Wiley & Sons, pp. 457-490.
Bernell, L. (1958) Water content and its effects on settlements in earth dams.
Proceedings of the 6th International Congress on Large Dams, New York, ICOLD.
pp. 373-384 (Q.22 R.117).
Bernell, L. (1982) Experiences of wet compacted dams in Sweden. Proceedings of the
14th International Congress on Large Dams, Rio de Janiero, ICOLD. pp. 421-431
(Q.55 R.24).
Berti, G., Villa, F., Dovera, D., Genevois, R. and Brauns, J. (1988) The disaster of
Stava/Northern Italy. Hydraulic Fill Structures, ASCE Speciality Conference,
Geotechnical Special Publication No. 21, (Van Zyl and Vick ed.) Colorado, ASCE,
New York. pp. 492-510.
Bertuccioli, P., D'Elia, B. and Esu, F. (1996) Instrumental prediction of a high cut
stability in a jointed o.c. clay. Proceedings of the Seventh International Symposium
on Landslides, (Senneset ed.) Trondheim, Norway, Balkema. Vol. 3, pp. 1509-
1514.
Bialostocki, R.J. (1961) Waipapa Power Project: Inspection report on field control
testing of earth dam materials. New Zealand Ministry of Works.
Binnie, G.M. (1978) The collapse of Dale Dyke dam in retrospect. Quarterly Journal of
Engineering Geology, Vol. 11, pp. 305-324.
Bishop, A.W. (1952) The stability of earth dams. Ph.D. Thesis, University of London.
References Page R5

Bishop, A.W. (1957) Some factors controlling the pore pressure set up during the
construction of earth dams. Proceedings of the 4th International Conference on Soil
Mechanics and Foundation Engineering, London, Butterworths Scientific
Publications, London. Vol. 2, pp. 294-300.
Bishop, A.W. (1967) Progressive failure - with special reference to the mechanism
causing it. Proceedings of the Geotechnical Conference on Shear Strength
Properties of Natural Soils and Rocks, Oslo, Vol. 2, pp. 142-150.
Bishop, A.W. (1973) The stability of spoil heaps. Quarterly Journal of Engineering
Geology, Vol 6, pp. 335-376.
Bishop, A.W., Hutchinson, J.N., Penman, A.D.M. and Evans, H.E. (1969) Geotechnical
investigation into the causes and circumstances of the disaster of 21 October 1966.
In A selection of technical reports submitted to the Aberfan Tribunal H.M.S.O.,
London, Welsh Office, pp. 1-80, (Item 1).
Bishop, A.W. and Vaughan, P.R. (1962) Selset Reservoir: design and performance of
the embankment. Proceedings of the Institution of Civil Engineers, Vol. 21 (Feb.),
pp. 305-345.
Bister, D., Fry, J.J., Costaz, J., Houis, J., Dupas, J.M., Degoutte, G., Lino, M. and
Rizzoli, J.L. (1994) Reassessment of earthfill and rockfill dams safety - case
histories. Proceedings of the 18th International Congress on Large Dams, Durban,
ICOLD. Vol. 1, pp. 645-670 (Q.68 R.43).
Bjerrum, L. (1967) Progressive failure in slopes of overconsolidated plastic clay and
clay shales (Terzaghi Lecture). A.S.C.E., Journal of the Soil Mechanics and
Foundations Division, Vol. 93 (SM5), pp. 2-49.
Bjerrum, L. (1972) Embankments on soft ground. Proceedings, ASCE Speciality
Conference on Performance of Earth and Earth Retaining Structures, Purdue
University, Lafayette, Vol. 2, pp. 1-54.
Bjerrum, L. (1973) Problems of soil mechanics and construction on soft clays and
structurally unstable soils (collapsible, expansive and others). Proceedings, Eighth
International Conference on Soil Mechanics and Foundation Engineering,
Moscow, Vol. 3, pp. 111-159.
Bjerrum, L., Loken, T., Heiberg, S. and Foster, R. (1969) A field study of factors
responsible for quick clay slides. Seventh International Congress on Soil
Mechanics & Foundation Engineering, Mexico, Vol. 1, pp. 531-540.
References Page R6

Bleasdale, A. (1969) Meteorological conditions relating to the Aberfan inquiry. In A


selection of technical reports submitted to the Aberfan Tribunal H.M.S.O., London,
Welsh Office, pp. 207-218 (Item 7).
Blee, C.E. and Meyer, A.A. (1955) Measurement of settlements at certain dams on the
TVA system and assumptions for earthquake loadings for dams in the TVA area.
Proceedings of the 5th International Congress on Large Dams, Paris, ICOLD. pp.
141-157 (Q.18 R.5).
Blee, C.E. and Reigel, R.M. (1951) Rock fill dams. Proceedings of the 4th International
Congress on Large Dams, New Delhi, ICOLD. pp. 189-208 (Q.13 R.22).
Blight, G. (1997) Destructive mudflows as a consequence of tailings dyke failures.
Geotechnical Engineering, Proc. Institution of Civil Engineers, Vol 125 (Jan), pp.
9-18.
Blight, G.E., Robinson, M.J. and Diering, J.A.C. (1981) The flow of slurry from a
breached tailings dam. Jnl of the South African Inst. of Mining and Metallurgy, Vol
81 (1), pp. 1-8.
Blinder, S., Toniatti, N.B. and Ribeiro, M.S. (1992) Construction progress and
behavioural monitoring of Segredo dam. Water Power & Dam Construction
(April), pp. 18-21.
Blondeau, F., Queyroi, D., Peignaud, M., Mieussens, C., Levillian, J.P. and Vogien, M.
(1977) Instrumentation du remblai experimental. Proceedings, International
Symposium on Soft Clay, (Brenner & Brand ed.) Bangkok, Asian Institute of
Technology. pp. 419-435.
Bodtman, W.L. and Wyatt, J.D. (1985) Design and performance of Shiroro rockfill
dam. Proceedings of the Symposium on Concrete Face Rockfill Dams - Design,
Construction and Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE
New York. pp. 231-251.
Borovoi, A.A. and Mikhailov, L.P. (1978) The Nurek multipurpose development. Water
Power & Dam Construction, Vol. 30 (9), pp. 53-55.
Borovoi, A.A., Mikhailov, L.P., Moiseev, I.S. and Radchenko, V.G. (1982) Soils for
and methods of embankment dam construction. Proceedings of the 14th
International Congress on Large Dams, Rio de Janiero, ICOLD. Vol. 4, pp. 503-
515 (Q.55 R.29).
Boucaut, W.R.P. and Beal, J.C. (1979) The engineering geology of the Little Para dam
site. Report Bk. 79/2, Department of Mines and Energy, South Australia.
References Page R7

Bowers, N.A. (1928) Uncompleted Lafayette rolled-fill earth dam damaged by


movement. Engineering News Record (27 Sept.), pp. 483-485.
Bowling, A.J. (1978) Mackintosh dam rockfill control testing. Report 4301-18, Hydro-
Electric Commission Tasmania, Civil Engineering Laboratories.
Bowling, A.J. (1979) Tullabardine dam rockfill control testing. Report 4333-1 and -2,
Hydro-Electric Commission Tasmania, Civil Engineering Laboratories.
Bowling, A.J. (1981) Laboratory investigations into the suitability of rockfill for
concrete faced rockfill dams. ANCOLD Bulletin, Vol. 59, pp. 21-29.
Bowling, A.J. (1981-82) Bastyan dam - rockfill control testing. Report 4395-1 to -6,
Hydro-Electric Commission Tasmania, Civil Engineering Laboratories.
Boyle, R.J. (1965) Benmore power station. Earth dam. Report on testing and
inspection. New Zealand Ministry of Works.
Brand, E.W. (1985a) Discussion: Embankment failure on clay near Rio de Janiero.
A.S.C.E., Journal of Geotechnical Engineering, Vol 11 (2), pp. 257-259.
Brand, E.W. (1985b) Predicting the failure of residual soil slopes. Proceedings of the
11th International Conference on Soil Mechanics and Foundation Engineering, San
Francisco, Vol. 5, pp. 2541-2578.
Brand, E.W. (1989) Correlation between rainfall and landslides. Proceedings 12th
International Conference on Soil Mechanics and Foundation Engineering, Rio,
Vol. 5, pp. 3091-3093.
Brand, E.W. (1995) Keynote paper: Slope instability in tropical areas. Proceedings of
the Sixth International Symposium on Landslides, (Bell ed.) Christchurch, Balkema,
Rotterdam. Vol. 3, pp. 2031-2051.
Brand, E.W. and Premchitt, J. (1989) Comparison of the predicted and observed
performance of the Muar test embankment. Proceedings, International Symposium
on Trial Embankments on Malaysian Marine Clays, (Hudson, Toh & Chan ed.)
Kuala Lumpur, Malaysian Highway Authority. Vol. 2, pp. 10.1-10.29.
Brand, E.W., Premchitt, J. and Phillipson, H.B. (1984) Relationship between rainfall
and landslides in Hong Kong. Proceedings of the Fourth International Symposium
on Landslides, Toronto, Canada, Balkema. Vol. 1, pp. 377-384.
Bravo, G. (1979) The excavation to support the core of Canales dam. Proceedings of the
13th International Congress on Large Dams, New Delhi, ICOLD. Vol. 1, pp. 1221-
1232 (Q.48 R.69).
References Page R8

Bravo, G., Giron, F. and Olalla, C. (1994) Behaviour models and actual behaviour of
Canales dam. Proceedings of the 18th International Congress on Large Dams ,
Durban, ICOLD. Vol. 3, pp. 293-306 (Q.70 R.20).
Breznik, M. (1979) The reliability of and damage to underground dams and other cut-
off structures in karst regions. Proceedings of the 13th International Congress on
Large Dams, New Delhi, ICOLD. Vol. 4, pp. 57-79 (C.04).
Brousek, M. (1976) Technology of placing the rock material at the rockfill dam of
Dalesice as to the exploitation of aggregates of worse quality. Proceedings of the
12th International Congress on Large Dams, Mexico, ICOLD. Vol. 1, pp. 697-707
(Q.44 R.36).
Bryson, V.R. (1980) Lake Pukaki dam - Earth dam construction. Ministry of Works
and Development, New Zealand.
Burland, J.B., Longworth, T.I. and Moore, J.F.A. (1977) A study of ground movement
and progressive failure caused by a deep excavation in Oxford Clay. Geotechnique,
Vol. 27 (4), pp. 557-591.
Byrd, T. and Middleboe, S. (1984) Weak ground cited as Carsington fails. New Civil
Engineer (14 June).
Byrne, P.M. and Beaty, M. (1998) "Post-liquefaction" theoretical/conceptual issues.
National Science Foundation Workshop on Shear Strength of Liquefied Soils,
(Stark Olson Kramer and Youd ed.) Urbana, Illinois, National Science Foundation.
pp. 10-39.
Caine, N. (1980) The rainfall intensity - duration control of shallow landslides and
debris flows. Geografiska Annaler, Vol 62 A (1-2), pp. 23-27.
Campbell, D.B. and Shaw, W.H. (1978) Performance of a waste rock dump on
moderately to steeply sloping foundations. Proceedings of the First International
Symposium on Stability in Coal Mining, (Brawner and Dorling ed.) Vancouver,
Miller Freeman. pp. 395-405.
Cannon, S.H. and Ellen, S.D. (1988) Chapter 3. Rainfall that resulted in abundant
debris-flow activity during the storm. In Landslides, floods, and marine effects of
the storm of January 3-5, 1982, in the San Francisco Bay region, California. U.S.
Geological Survey Professional Paper 1434 (Ellen & Wieczorek ed.), U.S.
Geological Survey, pp. 27-33.
Cary, A.S. (1958) Rockfill dams: Performance of Mud Mountain dam. A.S.C.E.,
Journal of the Power Division, Vol. 84 (PO4), pp. 1745-1 to 3.
References Page R9

Casagrande, A. (1965) Role of the "calculated risk" in earthwork and foundation


engineering (Terzaghi lecture). A.S.C.E., Journal of the Soil Mechanics and
Foundations Division, Vol 91 (SM4), pp. 1-40.
Casales, V.L. (1976) Discussion during technical session of Question 47. Proceedings
of the 12th International Congress on Large Dams, Mexico City, ICOLD. Vol. 5,
pp. 495-500.
Casinader, R. (1987) Discussion on paper: Kotmale dam and observations on CFRD.
A.S.C.E., Journal of Geotechnical Engineering, 113 (10), pp. 1198-1200.
Casinader, R. and Watt, R.E. (1985) Concrete face rockfill dams of the Winneke
project. Proceedings of the Symposium on Concrete Face Rockfill Dams - Design,
Construction and Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE
New York. pp. 140-162.
Castro, G. (1998) "Post-liquefaction" shear strength from case histories. National
Science Foundation Workshop on Shear Strength of Liquefied Soils, (Stark, Olson,
Kramer and Youd ed.) Urbana, Illinois, National Science Foundation. pp. 53-57.
Castro, G., Poulos, S.J. and Leathers, F.D. (1985) Re-examination of the slide of Lower
San Fernando Dam. A.S.C.E., Journal of Geotechnical Engineering, Vol 111 (9),
pp. 1093-1106.
Castro, G., Seed, R.B., Keller, T.O. and Seed, H.B. (1992) Steady-state strength
analysis of Lower San Fernando Dam slide. A.S.C.E., Journal of Geotechnical
Engineering, Vol 118 (3), pp. 406-427.
Catanach, R. and McDaniel, T.N. (1972) Lateral deformation of a dam embankment.
Performance of Earth and Earth Retaining Structures, Vol. 1 (Part 1), pp. 867-883.
Cavounidis, S. and Höeg, K. (1977) Consolidation during construction of earth dams.
A.S.C.E., Journal of the Geotechnical Engineering Division, Vol 103 (GT10), pp.
1055-1067.
Cetin, H. (2002) Personal communication with Mr. H. Cetin of Cukurova University,
Turkey.
Cetin, H., Laman, M. and Ertunc, A. (2000) Settlement and slaking problems in the
world's fourth largest rock-fill dam, the Ataturk dam in Turkey. Engineering
Geology, Vol. 56, pp. 225-242.
Champa, S. and Mahatharadol, B. (1982) Construction of Srinagarind dam. Proceedings
of the 14th International Congress on Large Dams, Rio de Janiero, ICOLD. pp.
255-278 (Q.55 R.15).
References Page R10

Chan, Y.C., Pun, W.K., Wong, H.N., Li, A.C.O. and Yeo, K.C. (1996) Investigation of
some major slope failures between 1992 and 1995. GEO Report 52, Geotechnical
Engineering Office, Civil Engineering Department, Hong Kong.
Chandler, R.J. (1972) Lias clay: weathering processes and their effect on shear strength.
Geotechnique, Vol. 22 (3), pp. 403-431.
Chandler, R.J. (1974) Lias clay: the long-term stability of cutting slopes. Geotechnique,
Vol. 24 (1), pp. 21-38.
Chandler, R.J. (1976) The history and stability of two Lias clay slopes in the upper
Gwash valley, Rutland. Phil. Trans. R. Soc. Lond. A., 283, pp. 463-491.
Chandler, R.J. (1984a) Delayed failure and observed strengths of first time failures in
stiff clays: a review. Proceedings of the Fourth International Symposium on
Landslides, Toronto, Vol. 2, pp. 19-25.
Chandler, R.J. (1984b) Recent European experience of landslides in over-consolidated
clays and soft rocks. Proceedings of the Fourth International Symposium on
Landslides, Toronto, Vol. 2, pp. 61-81.
Chandler, R.J. and Tosatti, G. (1995) The Stava tailings dam failure, Italy, July 1985.
Proc. Institution of Civil Engineers: Geotechnical Engineering, Vol 113 (1), pp.
67-79.
Charles, J.A. (1976) The use of one-dimensional compression tests and elastic theory in
predicting deformations of rockfill embankments. Canadian Geotechnical Journal,
Vol. 13 (3), pp. 189-200.
Charles, J.A. (1986) The significance of problems and remedial works at British earth
dams. Proceedings of the BNCOLD-IWES Conference on Reservoirs, pp. 123-141.
Charles, J.A. (1997) Special problems associated with earthfill dams. Proceedings of the
19th International Congress on Large Dams, Florence, ICOLD. Vol. 2, pp. 1083-
1196 (GR. Q73).
Charles, J.A. (1998) Lives of embankment dams: construction to old age. Dams and
Reservoirs, Vol 8 (Dec.), pp. 11-23.
Charles, J.A. and Boden, J.B. (1985) The failure of embankment dams in the United
Kingdom. Proceedings of the Symposium on Failures in earthworks, London,
Thomas Telford. pp. 181-202.
Charles, J.A. and Tedd, P. (1991) Long term performance and ageing of old
embankment dams in the United Kingdom. Proceedings of the 17th International
Congress on Large Dams, Vienna, ICOLD. pp. 461-473 (Q.65 R.25).
References Page R11

Charles, J.A. and Watts, K.S. (1987) The measurement and significance of horizontal
earth pressures in the puddle clay cores of old earth dams. Proceedings of the
Institution of Civil Engineers, Vol 82 (Feb), pp. 123-152.
Church, M. and Miles, M.J. (1987) Meteorological antecedents to debris flow in
southwestern British Columbia; some case studies. Reviews in Engineering
Geology, Geological Society of America, Vol 7, pp. 63-79.
Clausen, C.J.F. (1972) Measurement of pore water pressures, settlements and lateral
deformations at a test fill on soft clay brought to failure at Mastemyr, Oslo.
Technical Report 11, Norwegian Geotechnical Institute, Oslo.
Clements, R.P. (1984) Post-construction deformation of rockfill dams. A.S.C.E.,
Journal of Geotechnical Engineering, Vol. 110 (7), pp. 821-840.
Clough, R.W. and Woodward, R.J. (1967) Analysis of embankment stresses and
deformations. A.S.C.E., Journal of the Soil Mechanics and Foundations Division,
Vol. 93 (SM4), pp. 529-549.
Coffey & Hollingsworth (1971) Study of weak clays within the foundation. Maroon
dam. Project No. 4260B/2, for Queensland Irrigation and Water Supply
Commission.
Cole, B.A. (1974) Gordon River power development - Stage 1, Scotts Peak Dam, design
report. Report CDR 264, Hydro-Electric Commission Tasmania, Civil Design
Division.
Cole, B.A. and Fone, P.J.E. (1979) Repair of Scotts Peak dam, Tasmania. Proceedings
of the 13th International Congress on Large Dams, New Delhi, ICOLD. pp. 211-
231 (Q.49 R.15).
Cole, B.R. and Cummins, P.J. (1981) Behaviour of Dartmouth dam during construction.
Proceedings of the 10th International Conference on Soil Mechanics and
Foundation Engineering, Stockholm, Balkema, Rotterdam. Vol. 1, pp. 81-85.
Collingham, E. (1997) Hope Valley dam safety evaluation: Preliminary pitting
investigation. Report Ref. SA Water: 4810/96, South Australian Water
Corporation.
Collingham, E.B. and Newman, R.J. (1977) A report on the field work and laboratory
testing for the Design Branch study of the Happy Valley dam. South Australian
Government, Engineering and Water Supply Department, Design Branch, Soils and
Foundations Section.
Commonwealth Department of Works (1970) Corin dam: Design and construction
report. CDW, Canberra Branch.
References Page R12

Connell Wagner (1998a) Cardinia Reservoir: Surveillance graphs 26-27/05/98. for


Melbourne Water Corporation.
Connell Wagner (1998b) Greenvale Reservoir: Surveillance graphs 31/3/98 to 1/4/98.
for Melbourne Water Corporation.
Connell Wagner (1998c) Thomson reservoir: Surveillance graphs. Ref. V018.01, for
Melbourne Water Corporation.
Connell Wagner (1998d) Upper Yarra reservoir, main dam: Surveillance graphs. for
Melbourne Water Corporation.
Connell Wagner (1999) Yan Yean reservoir: Surveillance graphs. for Melbourne Water
Corporation.
Convery, D.J. (1977) Upper Waitaka power development. Pukaki earth dam. Earth
dam materials report. New Zealand Ministry of Works and Development.
Cooke, J.B. (1958) Rockfill dams: Wishon and Courtright concrete face dams. A.S.C.E.,
Journal of the Power Division, Vol. 84 (PO3), pp. 1746-1 to 33.
Cooke, J.B. (1984) Progress in rockfill dams (18th Terzaghi lecture). ASCE, Journal of
Geotechnical Engineering, Vol 110 (10), pp. 1383-1414.
Cooke, J.B. (1993) Rockfill and the rockfill dam. Proceedings of the International
Symposium on High Earth-Rockfill Dams, (Jiang, Zhang & Qin ed.) Beijing,
Cooke, J.B. (1997) The concrete face rockfill dam. Non-Soil Water Barriers for
Embankment Dams, 17th Annual USCOLD Lecture Series, San Diego, California,
USCOLD. pp. 117-132.
Cooke, J.B. (1999) The development of today's CFRD dam. Proceedings of the Second
Symposium on Concrete Face Rockfill Dams, Florianopolis, Brazil, Brazilian
Committee on Dams. pp. 3-11.
Cooke, J.B. and Sherard, J.L. (1987) Concrete-face rockfill dam: II. Design. ASCE,
Journal of Geotechnical Engineering, Vol 113 (10), pp. 1113-1133.
Cooke, J.B. and Strassburger, A.G. (1988) Section 6: Rockfill Dams. In Development of
Dam Engineering in the United States (Kollgaard & Chadwick ed.), Permagon
Press, pp. 885-1030.
Cooling, L.F. and Golder, H.Q. (1942) The analysis of the failure of an earth dam
during construction. Journal of the Institution of Civil Engineers (No. 1 (1942-43),
Nov.), pp. 38-55.
Cooper, B., Khalili, N. and Fell, R. (1997) Large deformations due to undrained strain
weakening slope instability at Hume Dam No. 1 embankment. Proceedings of the
References Page R13

19th International Congress on Large Dams, Florence, ICOLD. Vol. 2, pp. 797-
818 (Q.73 R.46).
Cooper, M.R., Bromhead, E.N., Petley, D.J. and Grant, D.I. (1998b) The Selborne
cutting stability experiment. Geotechnique, Vol. 48 (1), pp. 83-101.
Cornforth, D.H., Worth, E.G. and Wright, W.L. (1974) Observations and analysis of a
flow slide in sand fill. Proc. of the Symposium on Field Instrumentation in
Geotechnical Engineering, (British Geotechnical Society ed.) England,
Butterworths, London. pp. 136-151.
Corominas, J. (1996a) The angle of reach as a mobility index for small and large
landslides. Canadian Geotechnical Journal, Vol 33, pp. 260-271.
Corominas, J. (1996b) Debris slide. In Landslide Recognition: Identification, Movement
and Causes (Dikau, Brunsden, Schrott and Ibsen ed.), John Wiley, pp. 97-102.
Corominas, J., Remondo, J., Farias, P., Estevao, M., Zezere, J., Diaz de Teran, J.,
Dikau, R., Schrott, L., Moya, J. and Gonzales, A. (1996) Debris flow. In Landslide
Recognition: Identification, Movement and Causes (Dikau, Brunsden, Schrott and
Ibsen ed.), John Wiley, pp. 161-180.
Costa-Filho, L.M., Werneck, M.L.G. and Collet, H.B. (1977) The undrained strength of
a very soft clay. Proceedings, Ninth International Symposium on Soil Mechanics
and Foundation Engineering, Tokyo, Vol. 1, pp. 69-72.
Coumoulos, D.G. (1979) Discussion: Engineering properties and performance of clay
fills. Proceedings ICE Conference on Clay Fills, London, The Institution of Civil
Engineers, London. pp. 221-224.
Coumoulos, D.G. and Koryalos, T.P. (1979) Performance of the clay core of a large
embankment dam during construction. Proceedings ICE Conference on Clay Fills,
London, The Institution of Civil Engineers, London. pp. 73-78.
Cox, F. (1972) The geology of the Lake William Hovell dam: First Stage. State Rivers
and Water Supply Commission Victoria.
Cribben, N.E. (1990) K.R.P.D., Crotty Dam, Field dry density testing on Zone 3A
material. Report 4652-1, Hydro-Electric Commission Tasmania, Engineering and
Scientific Services Department.
Crooks, J.H.A. (1987) Some observations on the stability of structures founded on soft
clays. Proceedings, International Symposium on Prediction and Performance in
Geotechnical Engineering, (Joshi & Griffiths ed.) Calgary, Balkema, Rotterdam.
pp. 27-38.
References Page R14

Crooks, J.H.A. and Becker, D.E. (1988) Discussion: Slide in the upstream slope of Lake
Shelbyville Dam by Humphreys and Leonards. Journal of Geotechnical
Engineering, ASCE, Vol. 114 (4), pp. 506-508.
Crooks, J.H.A., Becker, D.E., Jefferies, M.G. and McKenzie, K. (1984) Yield behaviour
and consolidation. I: Pore pressure response. pp. 356-381.
Cubrinovski, M. and Ishihara, K. (2000) Flow potential of sandy soils with different
grain compositions. Soils and Foundations, Vol 40 (4), pp. 103-119.
Currey, D.T., Michels, V. and Little, D.J. (1968) Bellfield dam, Victoria. Institution of
Engineers, Australia: Annual Engineering Conference Papers, IEAust, Sydney.
pp. 33-49.
Cyganiewicz, J.M. and Dise, K.M. (1997) Case history of a rapid drawdown failure at
Steinaker Dam. Proceedings of the 19th International Congress on Large Dams,
Florence, ICOLD. Vol. 4, pp. 481-497 (Q75 R36).
D'Appolonia, D.J., Lambe, T.W. and Poulos, H.G. (1971a) Evaluation of pore pressures
beneath an embankment. A.S.C.E., Journal of the Soil Mechanics and Foundations
Division, Vol 97 (SM6), pp. 881-897.
D'Elia, B., Picarelli, L., Leroueil, S. and Vaunet, J. (1998) Geotechnical characterisation
of slope movements in structurally complex clay soils and stiff jointed clays. Italian
Geotechnical Engineer, Anno XXXII (No. 3), pp. 5-32.
Daehn, W.W. (1955) Behaviour of a rolled earth dam constructed on a compressible
foundation. Proceedings of the 5th International Congress on Large Dams, Paris,
ICOLD. pp. 171-191 (Q.18 R.7).
Dascal, O. (1987) Postconstruction deformations of rockfill dams. A.S.C.E., Journal of
Geotechnical Engineering, Vol. 113 (1), pp. 46-59.
Dascal, O. and Tournier, J.P. (1975) Embankments on soft and sensitive clay
foundation. A.S.C.E., Journal of the Geotechnical Engineering Division, Vol 101
(GT3), pp. 297-314.
Davidson, R.R., Vreugdenhil, R. and Foster, M. (2001) The dam safety upgrade at Lake
Eppalock. ANCOLD Bulletin, Issue 118 (August), pp. 59-70.
Davies, D.G. (1953) The Harrogate Dam failure. Journal of the Institution of Water
Engineers, Vol. 7 (1), pp. 57-79.
Davies, T.J.G. (1993) King River power development, Crotty Dam, Record Design
Report. Report CDR 614, Hydro-Electric Commission Tasmania, Consulting
Business Unit.
References Page R15

Davies, T.J.G., Smith, O. and Hancock, D.J. (1995) Reece dam, Surveillance Report No.
2 (1988 - 1995). Report ENE-0010-SF-007, Hydro-Electric Commission
Tasmania, Civil and Water Resources Engineering Group.
Davis, F.J. and Kisselman, H.E. (1963) Summary of earthworks control for Navajo dam.
Upper Colorado River Storage Project. United States Department of the Interior
Bureau of Reclamation, Office of Chief Engineer.
Dawson, R.F. (1994) Mine waste geotechnics. Doctor of Philosophy, Department of
Civil Engineering, University of Alberta, Edmonton.
Dawson, R.F., Morgenstern, N.R. and Stokes, A.W. (1998) Liquefaction flowslides in
Rocky Mountain coal mine waste dumps. Canadian Geotechnical Journal, Vol 35,
pp. 328-343.
de Alba, P. (1998) Written statement: Shear strength of liquefied soils from laboratory
and field tests. Discussion. National Science Foundation Workshop on Shear
Strength of Liquefied Soils, (Stark Olson Kramer and Youd ed.) Urbana, Illinois,
National Science Foundation. pp. 113-115.
de Groot, M.B. and Stoutjesdijk, T.P. (1997) Undrained stress path of loose sand
predicted from dry tests. Canadian Geotechnical Journal, Vol 34, pp. 131-138.
de Melo, F.G. and Direito, F.T. (1982) The behaviour of Roxo dam. Proceedings of the
14th International Congress on Large Dams, Rio de Janiero, ICOLD. pp. 387-400
(Q.52 R.23).
Department of Public Works New South Wales (1955) Schedule of rate contract for
construction of embankment, intake structure, stilling basin and certain other works
at Adaminaby dam. Contract No. 17-55/56.
Department of Water Resources New South Wales (1989) Chaffey dam - design report.
Public Works Department, Dams Section.
Dibiagio, E., Myrvoll, F., Valstad, T. and Hansteen, H. (1982) Field instrumentation,
observations and performance of Svartevann dam. Proceedings of the 14th
International Congress on Large Dams, Rio de Janiero, ICOLD. pp. 789-826 (Q.52
R.49).
Dighe, M.R., Gaikwad, R.S., Nemade, V.D. and Jawalge, M.M. (1985) Failure analysis
of Aran Dam. Fifteenth International Congress on Large Dams, Lausanne, ICOLD.
pp. 1117-1124 (Q.57 C.4).
Dixon, H.H. (1958) Moisture control and compaction methods used during construction
of the Sasumua dam, Kenya. Proceedings of the 6th International Congress on
Large Dams, New York, ICOLD. pp. 139-152 (Q.22 R.10).
References Page R16

Dobry, R. and Alvarez, L. (1967) Seismic failure of Chilean tailings dams. A.S.C.E.,
Journal of the Soil Mechanics and Foundations Division, Vol 93 (SM6), pp. 237-
260.
Dolezalova, M. and Leitner, F. (1981) Prediction of Dalesice dam performance.
Proceedings of the 10th International Conference on Soil Mechanics and
Foundation Engineering, Stockholm, Balkema, Rotterdam. Vol. 1, pp. 111-114.
Douglas, A.W. (1965) Burrendong dam construction. I.E. Aust., Civil Engineering
Transactions (October), pp. 49-62.
Dounias, G.T., Potts, D.M. and Vaughan, P.R. (1988) Finite element analysis of
progressive failure: two case studies. Computers and Geotechnics. Special issue on
Embankment Dams, Vol. 6 (2), pp. 155-175.
Dounias, G.T., Potts, D.M. and Vaughan, P.R. (1996) Analysis of progressive failure
and cracking in old British dams. Geotechnique, Vol 46 (4), pp. 621-640.
Downer Energy Services Ltd (1998) Waipapa dam: Report on the Type B deformations
survey. for Electricity Corporation of New Zealand, Northern Generation Group.
Duncan, J.M. (1996) State of the art: Limit equilibrium and finite-element analysis of
slopes. A.S.C.E., Journal of Geotechnical Engineering, Vol 122 (7), pp. 577-595.
Duncan, J.M. and Dunlop, P. (1969) Slopes in stiff fissured clays and shales. A.S.C.E.,
Journal of the Soil Mechanics and Foundations Division, Vol. 95 (SM2), pp. 467-
492.
Duncan, J.M., Lefebvre, G. and Lade, P. (1980) The landslide at Tuve, near Goteborg,
Sweden, on November 30, 1977. Committee on Natural Disasters, Commission on
Sociotechnical Systems, National Research Council.
Duncan, J.M., Wright, S.G. and Wong, K.S. (1990) Slope stability during rapid
drawdown. Proceedings, H. Bolton Seed Memorial Symposium, (Duncan ed.)
BiTech, Vancouver. Vol. 2, pp. 253-272.
Eadie, A. (1988) Geological investigations for construction of The Bjelke-Petersen dam.
Ref. No. PP 1700, Queensland Water Resources Commission, Planning Division,
Engineering Geology Section.
Eckersley, J.D. (1985) Flowslides in stockpile coal. Engineering Geology, Vol 22, pp.
13-22.
Eckersley, J.D. (1986) The initiation and development of slope failures with particular
reference to flowslides. Doctor of Philosophy, James Cook University of North
Queensland, Department of Civil & Systems Engineering.
References Page R17

Eckersley, J.D. (1990) Instrumented laboratory flowslides. Geotechnique, Vol 40 (3),


pp. 489-502.
Eden, W.J. (1977) Evidence of creep in steep natural slopes of Champlain sea clay.
Canadian Geotechnical Journal, Vol 14, pp. 620-627.
Eden, W.J., Fletcher, E.B. and Mitchell, R.J. (1971) South Nation River landslide, 16
May 1971. Canadian Geotechnical Journal, Vol 8, pp. 446-451.
Edgars, L. and Karlsrud, K. (1983) Soil flows generated by submarine slides - Case
studies and consequences. Proc. of the Third International Conference on the
Behaviour of Off-shore Structures, (Connor and Chryssostomos ed.) MIT,
Hemisphere Pub. Corp., pp. 425-437.
Eide, O. and Bjerrum, L. (1956) The slide at Bekkelaget. Geotechnique, Vol 6, pp. 88-
100.
Eigenheer, L.P.Q.T. and Souza, R.J.B. (1999) Xingo concrete face rockfill dam.
Proceedings of the Second Symposium on Concrete Face Rockfill Dams,
Florianopolis, Brazil, Brazilian Committee on Dams. pp. 271-284.
Eisenstein, Z. and Law, S.T.C. (1977) Analysis of consolidation behaviour of Mica
dam. A.S.C.E., Journal of the Geotechnical Engineering Division, Vol. 103 (GT8),
pp. 879-895.
El-Ramley, H. (2001) Probabilistic analysis of landslides hazards and risks: Bridging
theory and practice. Ph.D. Thesis, Department of Civil and Environmental
Engineering, University of Alberta, Edmonton.
Ellen, S.D., Algus, M.A., Cannon, S.H., Fleming, R.W., Lahr, P.C., Peterson, D.M. and
Reneau, S.L. (1988) Chapter 6. Description and mechanics of soil slip/debris flows
in the storm. In Landslides, floods, and marine effects of the storm of January 3-5,
1982, in the San Francisco Bay region, California. U.S. Geological Survey
Professional Paper 1434 (Ellen & Wieczorek ed.), U.S. Geological Survey, pp. 63-
111.
Ellen, S.D. and Fleming, R.W. (1987) Mobilisation of debris flows from soil slips, San
Francisco Bay region, California. Reviews in Engineering Geology, Geological
Society of America, Vol 7, pp. 31-40.
Engineering and Water Supply Department (1981) Kangaroo Creek dam. Historical
account of Construction and Operations. Ref. 81/41, South Australian
Government.
References Page R18

Engineering and Water Supply Department South Australia (1981a) Happy Valley dam:
Historical account of construction and operations. Ref. No. 81/11, Design
Services Branch.
Engineering and Water Supply Department South Australia (1981b) Hope Valley dam:
Historical account of construction and operations. Ref. No. 81/10, Design
Services Branch.
ENR (1929a) Modified completion Lafayette Dam recommended. Engineering News
Record (17 Jan.), pp. 116.
ENR (1929b) Plastic foundations. Engineering News Record (31 January), pp. 167.
ENR (1929c) Reconstruction of Lafayette Dam advised. Engineering News Record (31
Jan.), pp. 190-192.
ENR (1937a) Foundation of earth dam fails. Engineering News Record (30 Sept.), pp.
532 & 535.
ENR (1937b) Small earthfill dam fails. Engineering News Record (24 June), pp. 932.
ENR (1937c) WPA dam fails at Kansas City. Engineering News Record (23 Sept.), pp.
495.
ENR (1938a) Foundation caused dam failure. Engineering News Record (24 Feb.), pp.
281-282.
ENR (1938b) Why Marshall dam failed. Engineering News Record (24 Mar.), pp. 431-
432.
ENR (1939a) Large slide in Fort Peck Dam caused by foundation failure. Engineering
News Record (11 May) pp. 55-58.
ENR (1939b) Small earthfill dam typical of modern practice. Engineering News Record
(26 Oct.), pp. 47.
ENR (1958) Near Denver, a 60 ft earth dam sunk 11 ft. Engineering News Record (26
June), pp. 23.
ENR (1960) High earth dam plugs narrow canyon for power. Engineering News Record
(7 Apr.), pp. 44-51.
Esmiol, E.E. (1955) Foundation consolidation beneath four Bureau of Reclamation
rolled earth dams as determined from field observations. Proceedings of the 5th
International Congress on Large Dams, Paris, ICOLD. pp. 123-139 (Q.18 R.3).
Evans, M.D. and Zhou, S. (1995) Liquefaction behavior of sand-gravel composites.
A.S.C.E., Journal of Geotechnical Engineering, Vol 121 (3), pp. 287-298.
References Page R19

Evans, N.C., Huang, S.W. and King, J.P. (1997) The natural terrain landslide study,
Phases I and II. Special Project Report SPR 5/97, Geotechnical Engineering
Office, Civil Engineering Department, Hong Kong.
Fannin, R.J. and Rollerson, T.P. (1992) Debris flows: some physical characteristics and
behaviour. Canadian Geotechnical Journal, Vol 30, pp. 71-81.
Fannin, R.J., Wise, M.P., Wilkinson, J.M. and Rollerson, T.P. (1996) Landslide
initiation and runout on clearcut hillslopes. Proceedings of the Seventh
International Symposium on Landslides, (Senneset ed.) Trondheim, Norway,
Balkema. Vol. 1, pp. 195-199.
Farhi, F.J. and Hamon, M. (1967) Djatiluhur dam. Fill deformations and core cracking
(in French). Proceedings of the 9th International Congress on Large Dams,
Istanbul, ICOLD. pp. 457-478 (C.10).
Fear, C.E. and McRoberts, E.C. (1995) Reconsideration of initiation of liquefaction in
sandy soils. A.S.C.E., Journal of Geotechnical Engineering, Vol 121 (3), pp. 249-
261.
Fell, R. (2002) Personnel communication with Prof. Fell of University of New South
Wales, Australia.
Fell, R., Hungr, O., Leroueil, S. and Riemer, W. (2000) Keynote lecture - Geotechnical
engineering of the stability of natural slopes, and cuts and fills in soil. Proceedings
of the International Conference on Geotechnical and Geological Engineering
(GeoEng2000), Melbourne, Technomic, Lancaster. Vol. 1, pp. 21-120.
Fell, R., MacGregor, J.P. and Stapleton, D. (1992) Geotechnical Engineering of
Embankment Dams. Balkema, Rotterdam, pp. 675.
Fell, R., MacGregor, J.P. and Stapleton, D. (in press) Geotechnical Engineering of
Embankment Dams.
Fell, R., Wong, P. and Stone, P. (1987) Slope instability in soft ground. Soil Slope
Instability and Stabilisation, (Walker & Fell ed.) Balkema, Rotterdam. pp. 231-278.
Ferkh, Z. and Fell, R. (1994) Design of embankments on soft clay. Proceedings, 13th
International Conference on Soil Mechanics and Foundation Engineering, New
Delhi, Balkema, Rotterdam. pp. 733-738.
Fetzer, C.A. (1967) Electro-osmotic stabilisation of West Branch Dam. A.S.C.E.,
Journal of the Soil Mechanics and Foundations Division, Vol. 93 (SM4), pp. 85-
106.
References Page R20

Finlay, P.J., Fell, R. and Maguire, P.K. (1997) Relationship between the probability of
landslide occurrence and rainfall. Canadian Geotechnical Journal, Vol. 34 (6), pp.
811-824.
Finlay, P.J., Mostyn, G.R. and Fell, R. (1999) Landslide risk assessment: prediction of
travel distance. Canadian Geotechnical Journal, Vol. 36, pp. 556-562.
Fitzpatrick, M.D., Cole, B.A., Kinstler, F.L. and Knoop, B.P. (1985) Design of
concrete-faced rockfill dams. Proceedings of the Symposium on Concrete Face
Rockfill Dams - Design, Construction and Performance, (Cooke and Sherard ed.)
Detroit, Michigan, ASCE New York. pp. 410-434.
Fitzpatrick, M.D., Liggins, T.B. and Barnett, R.H.W. (1982) Ten years of surveillance
of Cethana dam. Proceedings of the 14th International Congress on Large Dams,
Rio de Janiero, ICOLD. pp. 847-866 (Q.52 R.51).
Fitzpatrick, M.D., Liggins, T.B., Lack, L.J. and Knoop, B.P. (1973) Instrumentation and
performance of Cethana dam. Proceedings of the 11th International Congress on
Large Dams, Madrid, ICOLD. pp. 145-164 (Q.42 R.9).
Fleming, R.W., Ellen, S.D. and Algus, M.A. (1989) Transformation of dilative and
contractive landslide debris into debris flows - An example from Marin County,
California. Engineering Geology, Vol 27, pp. 201-223.
Folkes, D.J. and Crooks, J.H.A. (1985) Effective stress paths and yielding in soft clays
below embankments. Canadian Geotechnical Journal, Vol 22, pp. 357-374.
Folkes, D.J. and Crooks, J.H.A. (1986) Reply: Effective stress paths and yielding in soft
clays below embankments. Canadian Geotechnical Journal, Vol 23, pp. 413.
Forza, S.J. and Hancock, D.J. (1995a) Cethana dam, Surveillance Report No. 2 (1982 -
1995). Report ENE-0010-SF-004, Hydro-Electric Commission Tasmania, Civil
and Water Resources Engineering Group.
Forza, S. and Hancock, D. (1995b) Tullabardine dam, Surveillance Report No. 2 (1986
- 1994). Report ENE-0010-SF-002, Hydro-Electric Commission Tasmania, Civil
and Water Resources Engineering.
Forza, S., Hancock, D. and Young, A.J.A. (1993) Rowallan dam: Surveillance Report
No. 2 (1985 - 1993). Report No. CDR 590, Hydro-Electric Commission, Tasmania.
Forza, S.J. and Hancock, D.J. (1993) Serpentine dam, Surveillance Report No. 2 (1980 -
1993). Report CDR 602, Hydro-Electric Commission Tasmania, Consulting
Business Unit.
References Page R21

Foster, M.A. (1999) The probability of failure of embankment dams by internal erosion
and piping. Ph.D. Thesis, The University of New South Wales, School of Civil and
Environmental Engineering.
Foster, M.A., Fell, R. and Spannagle, M. (2000) The statistics of embankment dam
failures and accidents. Canadian Geotechnical Journal, Vol. 37, pp. 1000-1024.
Fourie, A.B. (2000) Static liquefaction as the mechanism for flow failures of tailings
dams under non-seismic conditions. Proceedings of the International Conference
on Geotechnical and Geological Engineering (GeoEng2000), Melbourne, Vol. 1,
pp. 1263-1269.
Fox, S.D. and Fyfe, G.E. (2001) It's more fun when you do it yourself - constructing the
Lake Eppalock main embankment remedial works project. ANCOLD Bulletin, Issue
118 (August), pp. 71-79.
Franks, C.A.M. (1995) Engineering Geological Assessment of landslide at Milestone
14.5 on the Castle Peak Road. Geological Report GR 1/95, Geotechnical
Engineering Office, Civil Engineering Department, Hong Kong.
Franks, C.A.M. (1996) Study of rainfall induced landslides on natural slopes in the
vicinity of Tung Chung New Town, Lantau Island. Special Project Report SPR
4/96, Geotechnical Engineering Office, Civil Engineering Department, Hong Kong.
Fredlund, D.G. and Rahardjo, H. (1993) Soil mechanics for unsaturated soils. John
Wiley & Sons, New York, pp. 517.
Fugro Scott Wilson Joint Venture (1999a) Detailed study of the landslide at Fung Wong
Reservoir on 9 June 1998. LSR 5/99, Geotechnical Engineering Office, Civil
Engineering Department, Government of Hong Kong.
Fugro Scott Wilson Joint Venture (1999b) Detailed study of the landslide below Au Tau
Village Road, Tseung Kwan O on 9 June 1998. LSR 6/99, Geotechnical
Engineering Office, Civil Engineering Department, Government of Hong Kong.
Fugro Scott Wilson Joint Venture (1999c) Detailed study of the landslide The Outward
Bound School on 9 June 1998. LSR 7/99, Geotechnical Engineering Office, Civil
Engineering Department, Government of Hong Kong.
Fugro Scott Wilson Joint Venture (1999d) Detailed study of the landslides at Yue Sun
Garden, Wo Mei on 9 June 1998. LSR 8/99, Geotechnical Engineering Office,
Civil Engineering Department, Government of Hong Kong.
Galloway, J.D. (1939) The design of rock-fill dams. ASCE, Transactions, Vol. 104, pp.
1-24.
References Page R22

Galloway, J.H.H. (1970) Matahina Power Project: Earth dam design report. Report
No. 92/12/75/4/1, Ministry of Works New Zealand, Power Design Office.
Gamboa, J. and Benassini, A. (1967) Behavior of Netzahualcoyotl dam during
construction. A.S.C.E., Journal of the Soil Mechanics and Foundations Division,
Vol. 93 (SM4), pp. 211-229.
Geotechnical Engineering Office (1994) Report on the Kwun Lung Lau landslide of 23
July 1994. Volume 2, Civil Engineering Department, Hong Kong Government.
Geotechnical Engineering Office (1996a) Report on the Fei Tsui Road landslide of 13
August 1995. Volume 2, Civil Engineering Department, Hong Kong Government.
Geotechnical Engineering Office (1996b) Report on the Shum Wan Road landslide of 13
August 1995. Volume 2, Civil Engineering Department, Hong Kong Government.
Gerke, D., Forza, S.J. and Hancock, D.J. (1995) Murchison dam, Surveillance Report
No. 2 (1985 - 1995). Report ENE-0010-SF-005, Hydro-Electric Commission
Tasmania, Civil and Water Resources Engineering Group.
Gerke, D. and Hancock, D. (1995) Parangana dam: Surveillance Report No. 2 (1986 -
1994). Report No. ENE-0010-SF-001, Hydro-Electric Commission Tasmania,
Civil and Water Resources Engineering.
Gerke, D., Quinlan, P. and Hancock, D.J. (1993) Scotts Peak dam, Surveillance Report
No. 2 (1985 - 1993). Report CDR 570, Hydro-Electric Commission Tasmania,
Consulting Business Unit.
Giron, F.C. (1997) Collapse settlement in the Canales dam. Proceedings of the 19th
International Congress on Large Dams, Florence, ICOLD. pp. 197-203 (Q73,
Contribution 2).
Giudici, S., Herweynen, R. and Quinlan, P. (2000) Hydro-Electric Commission
experience in concrete faced rockfill dams - past, present and future. Proceedings of
the International Symposium on Concrete Faced Rockfill Dams, Beijing, ICOLD.
pp. 29-46.
Giuliani, F.L. and Pujol, A. (1985) Seepage and slide downstream slope Arroyito Dam,
Argentina remedial measures. Fifteenth International Congress on Large Dams,
Lausanne, ICOLD. pp. 43-50 (Q.59 R.4).
Glastonbury, J. and Fell, R. (2000) Report on the analysis of "rapid" natural rock slope
failures. UNICIV Report R-390, The University of New South Wales, School of
Civil and Environmental Engineering.
References Page R23

Glastonbury, J., Fell, R. and Mostyn, G. (2002) Report on the post-collapse behaviour
of debris from rock slope failures. UNICIV Report No. R-406, The University of
New South Wales, School of Civil and Environmental Engineering.
Golder Associates Limited (1994) Failure run-out characteristics of mine waste dumps
in mountainous terrain. Unpublished report to the Canadian Centre for Mineral
and Energy Technology, Edmonton. Prepared in association with O. Hungr
Geotechnical Research Ltd.
Golder Associates Limited (1995) Mined rock and overburden piles: Runout
characteristics of debris from dump failures in mountainous terrain. Stage 2:
Analysis, modelling and prediction. Interim Report British Columbia Mine Waste
Rock Pile Research Committee and CANMET, In conjunction with O. Hungr
Geotechnical Research Ltd.
Goldsmith, R.C.M. (1977) Geological report on construction of the Googong dam,
Queanbeyan River, New South Wales, 1975-1977. Report No. GOG.51,
Department of National Development, Bureau of Mineral Resources, Geology and
Geophysics.
Gonzales, F.V. and Mena, E.S. (1997) Aguamilpa dam behaviour. Non-Soil Water
Barriers for Embankment Dams, 17th Annual USCOLD Lecture Series, San Diego,
California, USCOLD. pp. 133-147.
Good, R.J. (1976) Kangaroo Creek dam. Use of a weak schist as rockfill for a concrete
faced rockfill dam. Proceedings of the 12th International Congress on Large Dams,
Mexico, ICOLD. pp. 645-665 (Q.44 R.33).
Good, R.J. (1981) Behaviour of concrete faced rolled rockfill dams in South Australia.
ANCOLD Bulletin, Vol 59, pp. 45-56.
Good, R.J., Bain, D.L.W. and Parsons, A.M. (1985) Weak rock in two rockfill dams.
Proceedings of the Symposium on Concrete Face Rockfill Dams - Design,
Construction and Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE
New York. pp. 40-72.
Gosden, G.D., Belland, G.J. and Parsons, A.M. (2002) Hope Valley dam safety
investigation and remedial works. ANCOLD Bulletin, Issue 120 (April), pp. 89-99.
Gosschalk, E.M. and Kulasinghe, A.N.S. (1985) Kotmale dam and observations on
CFRD. Proceedings of the Symposium on Concrete Face Rockfill Dams - Design,
Construction and Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE
New York. pp. 379-395.
References Page R24

Gosschalk, E.M. and Kulasinghe, A.N.S. (1987) Closure to discussion on paper:


Kotmale dam and observations on CFRD. A.S.C.E., Journal of Geotechnical
Engineering, 113 (10), pp. 1202-1208.
Goulburn-Murray Water (1999) Lake Dartmouth: Report on dam safety surveillance
and behaviour (June 1998 to September 1998).
Gould, J.P. (1953) The compressibility of rolled fill materials determined from field
observations. Proceedings of the Third International Conference on Soil Mechanics
and Foundation Engineering, Switzerland, Vol. 2, pp. 239-244.
Gould, J.P. (1954) Compression characteristics of rolled fill materials in earth dams.
Technical Memorandum No. 648, United States Department of the Interior, Bureau
of Reclamation.
Government of Hong Kong (1977) Report on the slope failures at Sau Mau Ping August
1976. Report 67675-7K-1/77.
Graham, J., Crooks, J.H.A. and Bell, A.L. (1983) Time effects on the stress-strain
behaviour of natural soft clays. Geotechnique, Vol 33 (3), pp. 327-340.
Gray, E.W. (1972) Construction pore pressure dissipation in earth dams. Proceedings of
the Speciality Conference on Performance of Earth and Earth Retaining Structures,
Purdue University, Indianna, ASCE, New York. Vol. III, pp. 295-314.
Gregersen, O. (1981) The quick clay slide at Rissa, Norway. Proceedings of the 10th
International Conference on Soil Mechanics and Foundation Engineering,
Stockholm, Balkema, Rotterdam. Vol. 3, pp. 421-426.
Gregersen, O. and Loken, T. (1979) The quick-clay slide at Baastad, Norway, 1974.
Engineering Geology, Vol 14, pp. 183-196.
Gregory, C.H. (1844) On railway cuttings and embankments; with an account of some
slips in the London clay; on the line of the London and Croydon Railway. Minutes
of the Proceedings of Civil Engineers, Vol. 3 (March 26), pp. 135-173.
Grimston, J.O. (1989) Surveillance report: Geehi dam 1966 - 1989. Snowy Mountains
Hydro-Electric Authority, Civil Group.
Gu, W.H., Morgenstern, N.R. and Robertson, P.K. (1993) Progressive failure of Lower
San Fernando Dam. A.S.C.E., Journal of Geotechnical Engineering, Vol 119 (2),
pp. 333-349.
Guadagno, F.M. (1991) Debris flows in the Campanian volcaniclastic soils.
Proceedings of the International Conference on Slope Stability, Isle of Wight,
Thomas Telford. pp. 125-130.
References Page R25

Gutierrez, M. (1998) Written statement: Theoretical/Conceptual Issues. Discussion.


National Science Foundation Workshop on Shear Strength of Liquefied Soils,
(Stark Olson Kramer and Youd ed.) Urbana, Illinois, National Science Foundation.
pp. 88-89.
Gutteridge Haskins and Davey Pty Ltd (1995a) Surveillance review report for Geehi
dam. for Snowy Mountains Hydro-Electric Authority.
Gutteridge Haskins and Davey Pty Ltd (1995b) Surveillance review report for Tooma
dam. for Snowy Mountains Hydro-Electric Authority.
H.M.S.O. (1967) Report of the Tribunal appointed to inquire into the disaster at
Aberfan on October 21st, 1966. Her Majesty's Stationery Office, London.
Hacelas, J.E., Ramirez, C.A. and Regalado, G. (1985) Construction and performance of
Salvajina dam. Proceedings of the Symposium on Concrete Face Rockfill Dams -
Design, Construction and Performance, (Cooke and Sherard ed.) Detroit,
Michigan, ASCE New York. pp. 286-315.
Hadgraft, R.G. (1984) Design report for main dam: Bjelke-Petersen dam. (Internal
report) Queensland Water Resources Commission.
Halcrow Asia Partnership Ltd. (1998a) Report on the Ching Cheung Road landslide of
3 August 1997. Geotechnical Engineering Office, Civil Engineering Department,
Government of Hong Kong.
Halcrow Asia Partnership Ltd. (1998b) Report on the landslide at Ten Thousand
Buddhas' Monastery on 2 July 1997. Geotechnical Engineering Office, Civil
Engineering Department, Government of Hong Kong.
Halcrow Asia Partnership Ltd. (1998c) Report on the landslides at Hut No. 26 Kau Wa
Keng Upper Village on 4 June 1997. Geotechnical Engineering Office, Civil
Engineering Department, Government of Hong Kong.
Halcrow Asia Partnership Ltd. (1999a) Investigation of some selected landslide
incidents in 1997 (Volume 2). GEO Report 88, Geotechnical Engineering Office,
Civil Engineering Department, Government of Hong Kong.
Halcrow Asia Partnership Ltd. (1999b) Investigation of some selected landslide
incidents in 1997 (Volume 3). GEO Report 89, Geotechnical Engineering Office,
Civil Engineering Department, Government of Hong Kong.
Halcrow Asia Partnership Ltd. (1999c) Investigation of some selected landslide
incidents in 1997 (Volume 4). GEO Report 90, Geotechnical Engineering Office,
Civil Engineering Department, Government of Hong Kong.
References Page R26

Halcrow Asia Partnership Ltd. (1999d) Investigation of some selected landslide


incidents in 1997 (Volume 5). GEO Report 91, Geotechnical Engineering Office,
Civil Engineering Department, Government of Hong Kong.
Halcrow Asia Partnership Ltd. (1999e) Investigation of some selected landslide
incidents in 1997 (Volume 6). GEO Report 92, Geotechnical Engineering Office,
Civil Engineering Department, Government of Hong Kong.
Harder, L.F. (1997) Application of the Becker penetration test for evaluating the
liquefaction potential of gravelly soils. Proceedings of the NCEER Workshop on
Evaluation of Liquefaction Resistance of Soils, (Youd and Idriss ed.) Salt Lake
City, Utah, National Centre for Earthquake Engineering Research. pp. 129-148.
Hazen, A. (1918) A study of the slip in the Calaveras Dam. Engineering News Record,
Vol 81, (26 (Dec 26)), pp. 1158-1164.
Hazen, A. and Metcalf, L. (1918) Middle section of upstream slide of Calaveras Dam
slips into reservoir. Engineering News Record, Vol 80, (14 (April 4)), pp. 679-681.
He, G. (2000) Technical study on crest overflow of concrete face rockfill dams.
Proceedings of the International Symposium on Concrete Faced Rockfill Dams,
Beijing, ICOLD. pp. 283-291.
HECEC Australia Pty Ltd (1999) Tullaroop dam. Report on comprehensive dam
inspection March 1999. for Goulburn-Murray Water.
Heim, A. (1932) Landslides and human lives (Bergsturz and Menchen leben). N.
Skermer, translator. Bi-Tech Publishers, Vancouver, pp. 196.
Heinrichs, P.W. (1996) Mangrove Creek dam embankment behaviour and surveillance.
ANCOLD Bulletin, No. 102 (April), pp. 22-34.
Heitlinger, D., Moffatt, T.S. and Little, D.J. (1965) Design and construction of
Eppalock earth and rockfill dam, and turbine-pumping station, on the Campaspe
River, Victoria. The Journal of the Institution of Engineers, Australia (Oct-Nov),
pp. 325-356.
Henkel, D.J. (1956) Discussion on paper by Watson: Earth movement affecting L.T.E.
railway in deep cutting east of Uxbridge. Proceedings of the Institution of Civil
Engineers, Vol. 5, Part 2 (27 March), pp. 320-323.
Henkel, D.J. (1957) Investigations of two long-term failures in London clay slopes at
Wood Green and Northolt. Proceedings of the Fourth International Conference on
Soil Mechanics and Foundation Engineering, London, Vol. 2, pp. 315-320.
References Page R27

Henkel, D.J. (1961) The shear strength of saturated remoulded clays. Proceedings of the
ASCE Research Conference on Shear Strength of Cohesive Soils, Boulder,
Colorado, pp. 533-554.
Hickox, R.W. and Murray, B.C. (1983) Structural behavior report: Belle Fourche Dam.
United States Bureau of Reclamation, Division of Dam Safety, Structural Behavior
Branch.
Hickox, R.W. and Murray, B.C. (1984) Structural behavior report: San Luis dam,
Central Valley Project, California, Mid-Pacific Region. United States Bureau of
Reclamation, Division of Dam Safety, Structural Behavior Branch, Embankment
Dam Instrumentation Section.
Hilf, J.W. (1948) Estimating construction pore pressures in rolled earth dams.
Proceedings, 2nd International Conference on Soil Mechanics and Foundation
Engineering, Rotterdam, Vol. 3, pp. 234-240.
Hilton, J.R., Gibson, E.J.R. and Pinkerton, I.L. (1974) Geehi dam project. Geehi dam:
Final design report. Technical Memorandum No. C.D. 226, Snowy Mountains
Hydro-Electric Authority, Civil Engineering Design Division.
HKIE Geotechnical Division (1998) Soil nails in loose fill. A preliminary study. Draft
Report Hong Kong Institution of Engineers, Geotechnical Division.
Höeg, K., Andersland, O.B. and Rolfsen, E.N. (1969) Undrained behaviour of quick
clay under load tests as Asrum. Geotechnique, Vol 19 (1), pp. 101-115.
Holomek, S. (1994) Safety evaluation methods of Dalesice dam. Proceedings of the
18th International Congress on Large Dams, Durban, ICOLD. Vol. 1, pp. 19-28
(Q.68 R.03).
Holton, I.R. (1992) In-service deformation of a puddle clay core dam. Dams and
Reservoirs, Vol 2 (1), pp. 12-18.
Holtz, R.D. and Holm, G. (1979) Test embankment on an organic silty clay.
Proceedings, Seventh European Conference on Soil Mechanics and Foundation
Engineering, Brighton, England, British Geotechnical Society. Vol. 3, pp. 79-86.
Holtz, R.D. and Kovac, W.D. (1981) An Introduction to Geotechnical Engineering. (N.
M. Newmark and W. J. Hall ed.), Prentice Hall, New Jersey, pp. 733.
Hong, S.W., Sohn, J.I., Bae, G.J., Ahn, S.R., Um, Y.S. and Park, E.Y. (1994) A case
study of rockfill dam: Stability evaluation and remedial treatment. Proceedings of
the 13th International Conference on Soil Mechanics and Foundation Engineering,
New Delhi, Balkema, Rotterdam. pp. 967-970.
References Page R28

Howard, S.R., Gardner, P.E.J., McConnel, A.D., Gibson, E.J.R. and Pinkerton, I.L.
(1974) Khancoban and Murray 2 projects. Khancoban dam and excavation for
Murray 2 power station. Final design report. Technical Memorandum No. C.D.
277, Snowy Mountains Hydro-Electric Authority, Civil Engineering Design
Division.
Howard, S.R., Hilton, J.I., Bell, G.J., Gibson, E.J.R. and Pinkerton, I.L. (1978) Tumut 3
project. Talbingo dam, spillway, headrace channel & pipeline inlet structure. Final
design report. Technical Memorandum No. C.D. 285, Snowy Mountains Hydro-
Electric Authority, Civil Engineering Design Division.
Howard, T.R., Baldwin, J.E. and Donley, H.F. (1988) Chapter 9. Landslides in Pacifica,
California, caused by the storm. In Landslides, floods, and marine effects of the
storm of January 3-5, 1982, in the San Francisco Bay region, California. U.S.
Geological Survey Professional Paper 1434 (Ellen & Wieczorek ed.), U.S.
Geological Survey, pp. 163-183.
Howley, I.A. (1971) King River Project: Report on rockfill in the main embankment.
State Rivers and Water Supply Commission Victoria.
Howson, G.W. (1939) Discussion on paper by Galloway: The design of rock-fill dams.
ASCE, Transactions, Vol. 104, pp. 42-45.
Hsü, K.J. (1975) Catastrophic debris streams (sturzstroms) generated by rockfalls.
Geological Society of America Bulletin, Vol 86, pp. 129-140.
Huang, Y., Peng, Z., Si, H. and Li, G. (1993) Monitoring and performance of Xibeikou
CFRD. Proceedings of the International Symposium on High Earth-Rockfill Dams,
(Jiang, Zhang & Qin ed.) Beijing, Vol. 1, pp. 475-482.
Humphrey, D.N. and Leonards, G.A. (1986) Slide in the upstream slope of Lake
Shelbyville Dam. Journal of Geotechnical Engineering, ASCE, Vol. 112 (5), pp.
564-577.
Humphrey, D.N. and Leonards, G.A. (1988) Closure: Slide in the upstream slope of
Lake Shelbyville Dam. Journal of Geotechnical Engineering, ASCE, Vol. 114 (4),
pp. 510-513.
Hungr Geotechnical Research Inc. (1998) Mobility of landslide debris in Hong Kong:
Pilot back analyses using a numerical model. Geotechnical Engineering Office,
Hong Kong.
Hungr, O. (1990) Mobility of Rock Avalanches. Report of the National Research
Institute for Earth Science and Diaster Prevention, Tsukuba, Japan, No. 46, pp. 11-
20.
References Page R29

Hungr, O. (1995) A model for the runout analysis of rapid flow slides, debris flows, and
avalanches. Canadian Geotechnical Journal, Vol 32, pp. 610-623.
Hungr, O., Dawson, R.F., Kent, A., Campbell, D. and Morgenstern, N.R. (1998) Rapid
flow slides of coal mine waste in British Columbia, Canada. Submitted for
Publication to Canadian Geotechnical Journal.
Hungr, O. and Kent, A. (1995) Coal mine waste dump failures in British Columbia,
Canada. Landslide News, No. 9 (Dec.), pp. 26-28.
Hunter, G. and Fell, R. (2001) "Rapid" failure of soil slopes. UNICIV Report No. R-
400, The University of New South Wales, School of Civil and Environmental
Engineering.
Hunter, G. and Fell, R. (2002a) The deformation behaviour of rockfill. UNICIV Report
No. R-405, The University of New South Wales, School of Civil and
Environmental Engineering.
Hunter, G. and Fell, R. (2002b) Post failure deformation of slides in embankment dams
and cut slopes in over-consolidated high plastic clays. UNICIV Report No. R-411,
The University of New South Wales, School of Civil and Environmental
Engineering.
Hunter, G. and Fell, R. (2003) The deformation behaviour of embankment dams.
UNICIV Report No. R-416, The University of New South Wales, School of Civil
and Environmental Engineering.
Hunter, G., Fell, R. and Khalili, N. (2000) The deformation behaviour of embankments
on soft ground. UNICIV Report No. R-391, The University of New South Wales,
School of Civil and Environmental Engineering.
Hunter, J.R. and Bacon, G.A. (1970) Behaviour of Blowering dam embankment.
Proceedings of 10th the International Congress on Large Dams, Montreal, ICOLD.
pp. 217-242 (Q.38 R.15).
Hunter, J.R., Gibson, E.J.R. and Pinkerton, I.L. (1974) Tooma - Tumut project. Tooma
dam: Final design report. Technical Memorandum No. C.D. 227, Snowy
Mountains Hydro-Electric Authority, Civil Engineering Design Division.
Hutchinson, J.N. (1961) A landslide on a thin layer of quick clay at Furre, central
Norway. Geotechnique, Vol 11 (2), pp. 69-94.
Hutchinson, J.N. (1965) The landslide of February, 1959, at Vibstad in Namdalen.
Publication 61, Norwegian Geotechnical Institute.
Hutchinson, J.N. (1986) A sliding-consolidation model for flow slides. Canadian
Geotechnical Journal, Vol 23, pp. 115-126.
References Page R30

Hutchinson, J.N. (1988) General Report: Morphological and geotechnical parameters of


landslides in relation to geology and hydrogeology. Proceedings of the Fifth
International Symposium on Landslides, (Bonnard ed.) Lausanne, Switzerland,
Balkema. Vol. 1, pp. 3-35.
Hydro Electric Commission (1991a) Dams Surveillance Report, Bastyan Dam, 1986 -
1991 (No. 2). Report CDR 560, HEC Tasmania, Safety of Dams Unit.
Hydro Electric Commission (1991b) Dams Surveillance Report, Mackintosh Dam, 1985
- 1991 (No. 2). Report CDR 559, HEC Tasmania, Safety of Dams Unit.
Hydro Electric Commission Tasmania (1987) Anthony power development,
specification for excavation and rockfill for White Spur dam. Specification C.E.
2023.
Hydro Technology (1995) Cairn Curran Reservoir: Safety Surveillance 1994 and
Formal Dam Safety Inspection 17 November 1994. Report No. 90/10351, Rural
Water Corporation, Victoria.
ICOLD (1974) Lessons from Dam Incidents. In (Committee on Failures and Accidents
to Large Dams ed.), ICOLD, pp. 1069.
ICOLD (1989a) Moraine as embankment and foundation material. Bulletin No. 69,
International Commission on Large Dams.
ICOLD (1989b) Rockfill dams with concrete facing. State of the art. Bulletin 70,
International Commission on Large dams.
ICOLD (1993) Rock materials for rockfill dams. Bulletin No. 92, International
Commission on Large Dams.
Ikeya, H. (1989) Debris flows and its countermeasures in Japan. Bulletin of the
International Association of Engineering Geology, Bulletin No. 40, pp. 15-33.
Imam, S.M.R., Morgenstern, N.R., Robertson, P.K. and Chan, D.H. (2002) Yielding
and flow liquefaction of loose sand. Soils and Foundations, Vol 42 (3), pp. 19-31.
Indraratna, B., Balasubramanian, A.S. and Balachandran, S. (1992) Performance of test
embankment constructed to failure to failure on soft marine clay. A.S.C.E., Journal
of Geotechnical Engineering, Vol 118 (1), pp. 12-33.
Irfan, T.Y. (1997) Occurrence investigation and analysis of structurally controlled
landslides in weathered rock. Submitted for Publication to Quarterly Journal
Engineering Geology.
Irfan, T.Y. and Woods, N.W. (1998) Discontinuity controlled landslides in weathered
rocks in Hong Kong. Submitted for Publication to Transactions, Hong Kong
Institution of Engineers.
References Page R31

Ishihara, K. (1985) Stability of natural deposits during earthquakes. Proceedings of the


11th International Conference on Soil Mechanics and Foundation Engineering, San
Francisco, Vol. 1 pp. 321-376.
Ishihara, K. (1993) Liquefaction and flow failure during earthquakes. Geotechnique,
Vol 43 (3), pp. 351-415.
IUGS Working Group on Landslides (1995) A suggested method for describing the rate
of movement of a landslides. Bulletin 52, pp 75-78, International Association of
Engineering Geology.
Iverson, R.M., Reid, M.E. and LaHusen, R.G. (1997) Debris flow mobilisation from
landslides. Annual Review of Earth and Planetary Sciences, Vol. 25, pp. 85-138.
Jacobson, G. (1965) Nillahcootie damsite: Geology. Progress Report No. 1, State
Rivers and Water Supply Commission, Victoria.
James, P.M. (1970) Time effects and progressive failure in clay slopes. Ph.D. Thesis,
London University.
Japanese National Committee on Large Dams (1976) Dams in Japan. JNCOLD.
Jardine, R.J. and Hight, D.W. (1987) The behaviour and analysis of embankments on
soft clay. In Embankments on Soft Ground Public Works Research Centre, Greece,
pp. 195-244.
Jefferies, M. (1998) Written statement: Theoretical/Conceptual Issues. Discussion.
National Science Foundation Workshop on Shear Strength of Liquefied Soils,
(Stark Olson Kramer and Youd ed.) Urbana, Illinois, National Science Foundation.
pp. 93-97.
Jennings, J.E. (1979) The failure of a slimes dam at Bafokeng, mechanism of failure and
associated design considerations. The Civil Engineer in South Africa, Vol 21 (6),
pp. 135-141.
Jeyapalan, J.K., Duncan, J.M. and Seed, H.B. (1983a) Analyses of flow failures of mine
tailings dams. A.S.C.E., Journal of the Geotechnical Engineering Division, Vol 109
(2), pp. 150-171.
Jeyapalan, J.K., Duncan, J.M. and Seed, H.B. (1983b) Investigation of flow failures of
tailings dams. A.S.C.E., Journal of the Geotechnical Engineering Division, Vol 109
(2), pp. 172-189.
Jiyuan, S., Zhu, B. and Liang, C. (2000) Characteristic and experience of the design,
construction and performance of TSQ-1 concrete faced rockfill dam. Proceedings
of the International Symposium on Concrete Faced Rockfill Dams, Beijing,
ICOLD. pp. 97-105.
References Page R32

Johnson, A.M. (1980) Disasters at Aberfan, South Wales and at Buffalo Creek, West
Virginia. National Academy of Sciences, Committee on Excess Spoils from Coal
Mining.
Johnson, A.M. and Rodine, J.R. (1984) Debris flow. In Slope Instability (Brunsden and
Prior ed.), John Wiley, pp. 257-361.
Jones, O.T. (1955) Construction of the Cobb earthfill dam. New Zealand Engineering
(15th November), pp. 353-360.
Jones, O.T. (1965) Design of Benmore earth dam. New Zealand Engineering (January),
pp. 13-23.
Josseaume, H., Blondeau, F. and Pilot, G. (1977) Etude du comportment non draine de
trios argyles molles application au calcul de remblais. Proceedings, International
Symposium on Soft Clay, (Brenner & Brand ed.) Bangkok, Asian Institute of
Technology. pp. 487-504.
Justo, J.L. (1991) Collapse: Its importance, fundamentals and modelling. In Advances in
Rockfill Structures (Maranha das Neves ed.), Kluwer Academic Publishers,
Netherlands, pp. 97-152.
Kanbayashi, Y., Maeoka, M. and Harita, K. (1979) Deformation at the interface
between embankment and foundation of Terauchi dam. Proceedings of the 13th
International Congress on Large Dams, New Delhi, ICOLD. pp. 459-479 (Q.48
R.26).
Karasawa, S., Shimazu, Y., Shirakawa, N. and Kuwashima, T. (1994) A consideration
of the behaviour of zoned rockfill dams. Proceedings of the 18th International
Congress on Large Dams, Durban, ICOLD. pp. 851-882 (Q.68 R.52).
Karlsrud, K. and Edgars, L. (1982) Some aspects of submarine slope stability. Proc.
NATO Workshop on Marine Slides and Other Mass Movements, (Nieuwenhuis,
Saxov and Svend ed.) Algarve, Portugal, Plenum Press. pp. 61-81.
Kawashima, T. and Kanazawa, K. (1982) Design of rockfill dams on weathered
foundation with large scale faults. Proceedings of the 14th International Congress
on Large Dams, Rio de Janiero, ICOLD. Vol. 2, pp. 75-99 (Q.53 R.5).
Kennard, J. (1955) Difficulty experienced in the construction of the Hollowell dam,
Northampton. Proceedings of the 5th International Congress on Large Dams, Paris,
ICOLD. pp. 467-474 (Q.18 R.58).
Kennard, J. and Kennard, M.F. (1962) Selset Reservoir: design and construction.
Proceedings of the Institution of Civil Engineers, Vol. 21 (Feb.), pp. 277-304.
References Page R33

Khalili, N. (2002) Personnel communication with Dr. Khalili of University of New


South Wales, Australia.
Khalili, N., Fell, R. and Tai, K.S. (1996) A simplified method for estimating failure
induced deformation in embankments. Proceedings, Seventh International
Symposium on Landslides, (Senneset ed.) Trondheim, Norway, Balkema,
Rotterdam. Vol. 2, pp. 1263-1268.
Kim, S.K., Hong, W.P. and Kim, Y.M. (1991) Prediction of rainfall triggered landslides
in Korea. Proceedings of the Sixth International Symposium on Landslides, (Bell
ed.) Christchurch, Balkema, Rotterdam. Vol. 2 pp. 989-994.
Kim, Y.I. (1979) Dam behaviour measured by embedded instruments. Proceedings of
the 13th International Congress on Large Dams, New Delhi, ICOLD. pp. 419-438
(Q.49 R.28).
King, J.P. (1996) The Tsing Shan debris flow. Special Project Report SPR 6/96
(Volumes 1, 2 and 3), Geotechnical Engineering Office, Civil Engineering
Department, Hong Kong.
King, J.P. (1997) Natural terrain landslide study, Damage to Lui Pok School by a
natural terrain landslide. Discussion Note DN 2/97, Geotechnical Engineering
Office, Civil Engineering Department, Hong Kong.
Kinney, J.L. and Bartholomew, C.L. (1987) Structural behaviour report: Pueblo dam.
United States Bureau of Reclamation, Division of Dam Safety, Structural Behavior
Branch, Concrete Dam Instrumentation Section.
Kisa, H. and Fukuroi, H. (1994) Safety evaluation of the deformation behaviour of
rockfill dams based on the long-term observation. Proceedings of the 18th
International Congress on Large Dams, Durban, ICOLD. pp. 829-850 (Q.68 R.51).
Kjœrnsli, B., Kvale, G., Lunde, J. and Mathiesen, J.B. (1982) Design, construction
control and perfermance of Svartevann earth-rockfill dam. Proceedings of the 14th
International Congress on Large Dams, Rio de Janiero, ICOLD. pp. 319-337 (Q.55
R.19).
Kleiner, D.E. (1976) Design and construction of an embankment dam to impound
gypsum wastes. 12th International Congress on Large Dams, Mexico City,
ICOLD. Vol. 1, pp. 235-249 (Q44 R12).
Knight, R.G. (1938) The subsidence of a rockfill dam and the remedial measures
employed at Eildon Reservoir, Australia. Journal of the Institution of Civil
Engineers Australia (March), pp. 111-208.
References Page R34

Knill, J. (1996a) Report on the Fei Tsui Road landslide of 13 August 1995. Volume 1,
for Geotechnical Engineering Office, Civil Engineering Department, Hong Kong
Government.
Knill, J. (1996b) Report on the Shum Wan Road landslide of 13 August 1995. Volume
1, for Geotechnical Engineering Office, Civil Engineering Department, Hong Kong
Government.
Knoop, B.P. (1982a) Performance design report on Murchison dam. Report CDR 433,
Hydro-Electric Commission Tasmania, Civil Design Section.
Knoop, B.P. (1982b) Report on the performance of Mackintosh Dam. Report CDR 421,
Hydro-Electric Commission Tasmania.
Knoop, B.P. and Lack, L.J. (1985) Instrumentation and performance of concrete faced
rockfill dams in the Pieman River power development. Proceedings of the 15th
International Congress on Large Dams, Lausanne, ICOLD. pp. 1103-1120 (Q.56
R.58).
Knox, G. (1927) Landslides in South Wales valleys. Proceedings of the South Wales
Institute of Engineers, Vol 43, pp. 161-247, 257-290.
Koerner, H.J. (1977) Flow mechanisms and resistances in the debris streams of rock
slides. Bulletin of the International Association of Engineering Geology, Vol. 16,
pp. 101-104.
Koppejan, A.W., Wamelan, B.M. and Weinberg, L.J.H. (1948) Coastal flow slides in
the Dutch province of Zeeland. Proceedings of the 2nd International Conference on
Soil Mechanics and Foundation Engineering, Rotterdam, Vol. 5, pp. 89-96.
Kotzias, P.C. and Stamatopoulos, A.C. (1975) Statistical quality control at Kastraki
earth dam. A.S.C.E., Journal of the Geotechnical Engineering Division, Vol. 101
(GT9), pp. 837-853.
Kovacevic, N. (1994) Numerical analyses of rockfill dams, cut slopes and road
embankments. Doctor of Philosophy, London University (Imperial College of
Science, Technology and Medicine), Faculty of Engineering.
Kovacevic, N., Potts, D.M., Vaughan, P.R., Charles, J.A. and Tedd, P. (1997) Assessing
the safety of old embankment dams by observing and analysing movement during
reservoir operation. Proceedings of the 19th International Congress on Large
Dams, Florence, ICOLD. Vol. 2, pp. 551-566 (Q73 R35).
Kramer, S.L. (1988) Triggering of liquefaction flow slides in coastal soil deposits.
Engineering Geology, Vol 26, pp. 17-31.
References Page R35

Kramer, S.L. (1998) Written statement: Theoretical/Conceptual Issues. Discussion.


National Science Foundation Workshop on Shear Strength of Liquefied Soils,
(Stark, Olson, Kramer and Youd ed.) Urbana, Illinois, National Science
Foundation. pp. 104-105.
Kuerbis, R., Negussey, D. and Vaid, Y.P. (1988) Effect of gradation and fines content
on the undrained response of sand. ASCE Speciality Conference on Hydraulic Fill
Structures, Geotechnical Special Publication No. 21, (Van Zyl and Vick ed.) Fort
Collins, Colorado, ASCE, New York. pp. 330-345.
Kulasinghe, A.N.S. and Tandon, G., N, (1993) Technical and behavioural aspects of
Kotmale dam. Proceedings of the International Symposium on High Earth-Rockfill
Dams, (Jiang, Zhang & Qin ed.) Beijing, pp. 233-244.
La Rochelle, P., Trak, B., Tavenas, F. and Roy, M. (1974) Failure of a test embankment
on a sensitive Champlain clay deposit. Canadian Geotechnical Journal, Vol 11, pp.
142-164.
Ladd, C.C. (1972) Test embankment on sensitive clay. Proceedings, ASCE Speciality
Conference on Performance of Earth and Earth Retaining Structures, Purdue
University, Lafayette, Vol. 1 (1), pp. 101-128.
Ladd, C.C. (1991) Stability evaluation during staged construction (The 22nd Karl
Terzaghi Lecture). ASCE, Journal of Geotechnical Engineering, Vol 117 (4), pp.
538-615.
Lade, P.V. (1992) Static instability and liquefaction of loose fine sandy slopes. A.S.C.E.,
Journal of Geotechnical Engineering, Vol 118 (1), pp. 51-71.
Lade, P.V. and Yamamuro, J.A. (1997) Effect of nonplastic fines on static liquefaction
of sands. Canadian Geotechnical Journal, Vol 34, pp. 918-928.
Lambe, T.W. (1973) Predictions in soil engineering. Geotechnique, Vol. 23 (2), pp.
149-202.
Lambe, T.W. and Whitman, R.V. (1979) Soil Mechanics. Wiley & Sons, New York,
pp. 553.
Land and Water Conservation NSW (1994) Wyangala dam surveillance report.
Land and Water Conservation NSW (1995a) Copeton dam surveillance report.
Land and Water Conservation NSW (1995b) Glenbawn dam surveillance report.
Land and Water Conservation NSW (1996) Blowering dam: Notes for Dam Safety
Committee Inspection.
Land and Water Conservation NSW (1997) Windemere dam surveillance report.
References Page R36

Lask, E. and Reinhardt, W.G. (1986) Turks building a rock mountain. Engineering
News Record, Vol. 217 (6), pp. 40-42.
Lauffer, H. and Schober, W. (1964) The Gepatsch rockfill dam in the Kauner Valley.
Proceedings of the 8th International Congress on Large Dams, Edinburgh, ICOLD.
pp. 635-660 (Q.34 R.39).
Lavallée, J.G., St-Arnaud, G., Gervais, R. and Hammamji, Y. (1992) Stability of the
Olga C test embankment. ASCE Geotechnical Special Publication No. 31, Stability
and Performance of Slopes and Embankments II, (Seed & Boulanger ed.) Berkeley,
California, ASCE, New York. Vol. 2, pp. 1006-1021.
Lawrence, D.E., Aylsworth, J.M. and Morey, C.R. (1996) Sensitive clay flows along the
South Nation River, Ontario, Canada. Proceedings of the Seventh International
Symposium on Landslides, (Senneset ed.) Trondheim, Norway, Balkema,
Rotterdam. Vol. 1, pp. 479-484.
Lefebvre, G. (1996) Soft sensitive clays. In Landslides - Investigation and Mitigation,
Transportation Research Board Special Report No. 247 (Turner and Schuster ed.),
National Academy Press, pp. 607-619.
Lefebvre, G., Rosenberg, P., Paquette, J. and Lavallee, J.G. (1991) The September 5,
1987, landslide on the La Grande River, James Bay, Quebec, Canada. Canadian
Geotechnical Journal, Vol 28 (2), pp. 263-275.
Legge, G.H.H. (1970) Mulungushi and Mita Hills dams - Operation and performance.
Proceedings of the 10th International Congress on Large Dams, Montreal, ICOLD.
pp. 71-90 (Q.38 R.6).
Leonard, G.K. and Raine, O.H. (1958) Rockfill dams: Performance of TVA central core
dams. A.S.C.E., Journal of the Power Division, Vol. 84 (PO4), pp. 1736-1 to 16.
Leroueil, S. (1996) Compressibility of clays: fundamental and practical aspects.
A.S.C.E., Journal of Geotechnical Engineering, Vol 122 (7), pp. 534-543.
Leroueil, S. (2001) Natural slopes and cuts: movement and failure mechanisms (Thirty-
ninth Rankine Lecture). Geotechnique, Vol. 51 (3), pp. 197-243.
Leroueil, S., Locat, J., Vaunat, J., Picarelli, L., Lee, H. and Faure, R. (1996)
Geotechnical characterisation of slope movements. Proceedings of the Seventh
International Symposium on Landslides, (Senneset ed.) Trondheim, Norway,
Balkema, Rotterdam. Vol. 1, pp. 53-74.
Leroueil, S. and Marques, M.E.S. (1996) Importance of strain rate and temperature
effects in geotechnical engineering. ASCE Geotechnical Special Publication No.
References Page R37

61, Measuring and Modeling Time Dependent Soil Behavior, (Sheahan & Kaliakin
ed.) Washington, D.C., ASCE, New York. pp. 1-60.
Leroueil, S. and Tavenas, F. (1986) Discussion: Effective stress paths and yielding in
soft clays below embankments. Canadian Geotechnical Journal, Vol 23, pp. 410-
413.
Leroueil, S., Tavenas, F. and Le Bihan, J.-P. (1983) Proprietes caracteristiques des
argiles de l'est du Canada. Canadian Geotechnical Journal, Vol. 20 (4), pp. 681-
705.
Leroueil, S., Tavenas, F., Mieussens, C. and Peignaud, M. (1978a) Construction pore
pressures in clay foundations under embankments. Part II: generalised behaviour.
Canadian Geotechnical Journal, Vol 16, pp. 66-82.
Leroueil, S., Tavenas, F., Trak, B., La Rochelle, P. and Roy, M. (1978b) Construction
pore pressures in clay foundations under embankments. Part I: the Saint-Alban test
fills. Canadian Geotechnical Journal, Vol 16, pp. 54-65.
Li, C.Y. (1963) Placing earthfill for Columbia's Troneras dam. Civil Engineering, Vol.
33 (February), pp. 37-39.
Li, C.Y. (1967) Construction pore pressures in three earth dams. A.S.C.E., Journal of
the Soil Mechanics and Foundations Division, Vol. 93 (SM2), pp. 1-26.
Li, S. (1991) Pieman power development, Reece dam, civil design report (record).
Report CDR 551, Hydro-Electric Commission Tasmania, Civil Design Section.
Li, S.S.Y., Giudici, S. and Tindall, W.F. (1993) Design of Crotty dam and spillway.
Proceedings of the International Symposium on High Earth-Rockfill Dams, Vol. 1,
pp. 263-271.
Li, S.S.Y., Paterson, S.J. and Kinstler, F.L. (1991) A concrete faced rockfill dam
constructed on a deeply weathered foundation. Proceedings of the 17th
International Congress on Large Dams, Vienna, ICOLD. pp. 1601-1612 (Q.66
R.85).
Li-Tianchi, C. (1983) A mathematical model for predicting the extent of a major
rockfall. Zietschrift fur Geomorphologie, Vol. 27 (4), pp. 473-482.
Liggins, T. (1971) Mersey-Forth power development, Cethana dam, design report.
Report CDR 218, Hydro-Electric Commission Tasmania, Civil Design Division.
Linell, K.A. and Shea, H.F. (1960) Strength and deformation characteristics of various
glacial tills in New England. ASCE, Research Conference on Shear Strength of
Cohesive Soils, University of Colorado, ASCE. pp. 275-314.
References Page R38

Lino, M. (1997) Le glissment dans le talus amont du barrages homogene en argile


humide de Mondely. Nineteenth International Congress on Large Dams, Florence,
ICOLD. Vol. 5, pp. 617-622 (Q.75-19).
List, F. and Beier, H. (1985) The Frauenau dam monitoring and observations.
Proceedings of the 15th International Congress on Large Dams, Lausanne,
ICOLD. pp. 335-351 (Q.56 R.17).
Lloyd, H.E., Moore, O.L. and Getts, W.F. (1958) Rockfill dams: Cherry Valley central
core dam. A.S.C.E., Journal of the Power Division, Vol. 84 (PO4), pp. 1733-1 to
24.
Lo, K.Y. and Lee, C.F. (1973) Analysis of progressive failure in clay slopes.
Proceedings of the Eighth International Conference on Soil Mechanics and
Foundation Engineering, Moscow, Vol. 1.1, pp. 251-258.
Locat, J. and Leroueil, S. (1997) Landslide stages and risk assessment issues in sensitive
clays and other soft sediments. Proc. International Workshop on Landslide Risk
Assessment, (Cruden and Fell ed.) Hawaii, Balkema, Rotterdam. pp. 261-270.
Macedo, G.G. (1999) Concrete face behaviour of Aguamilpa dam. Proceedings of the
Second Symposium on Concrete Face Rockfill Dams, Florianopolis, Brazil,
Brazilian Committee on Dams. pp. 211-222.
Macedo, G.G., Castro, A.J. and Montanez, C.L. (2000) Behaviour of Aguamilpa dam.
In Concrete Face Rockfill Dams, J. Barry Cooke Volume (Mori, Sobrinho, Dijkstra,
Guocheng and Borgatti ed.), Beijing, pp. 117-151.
Mackenzie, P.R. and McDonald, L.A. (1985) Mangrove Creek dam: Use of soft rock for
rockfill. Proceedings of the Symposium on Concrete Face Rockfill Dams - Design,
Construction and Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE
New York. pp. 208-230.
Magnan, J.P., Humbert, P. and Mouratidis, A. (1982) Finite element analysis of soil
deformations with time under an experimental embankment at failure. Proceedings,
International Symposium on Numerical Models in Geomechanics, (Dungar, Pande
& Studer ed.) Zurich, Balkema, Rotterdam. pp. 601-609.
Mahasandana, T. and Mahatraradol, B. (1985) Monitoring systems of Khao Laem dam.
Proceedings of the 15th International Congress on Large Dams, Lausanne,
ICOLD. pp. 1- (Q.56 R.1).
Malaysian Highway Authority (1989a) Factual report on the performance of the 13 trial
embankments. Proceedings, International Symposium on Trial Embankments on
References Page R39

Malaysian Marine Clays, (Hudson, Toh & Chan ed.) Kuala Lumpur, Malaysian
Highway Authority. Vol. 1.
Malaysian Highway Authority (1989b) 3 m high control embankment on untreated soft
ground. Proceedings, International Symposium on Trial Embankments on
Malaysian Marine Clays, (Hudson, Toh & Chan ed.) Kuala Lumpur, Malaysian
Highway Authority. Vol. 2, pp. 2.1-2.30.
Mann, M.J. and Snow, R.E. (1992) Analysis of slope failure and remedial design of an
earth dam. ASCE Geotechnical Special Publication No. 31, (Seed and Boulanger
ed.) Berkeley, California, ASCE. Vol. 2, pp. 923-939.
Marachi, N.D., Chan, C.K., Seed, H.B. and Duncan, J.M. (1969) Strength and
deformation characteristics of rockfill materials. Report No. TE-69-5, University
of California, Department of Civil Engineering.
Maranha das Neves, E., Ramos, C.M., Veiga Pinto, A. and Direito, M.T. (1994) Safety
improvement of Beliche dam. Proceedings of the 18th International Congress on
Large Dams, Durban, ICOLD. pp. 1167-1179 (Q.68 R.67).
Marche, R. and Chapuis, R. (1974) Controle de la stabilite des remblais par la mesure
des deplacements horizontaux. Canadian Geotechnical Journal, Vol 11, pp. 182-
201.
Marcuson, W.F., Ballard, R.F. and Ledbetter, R.H. (1979) Liquefaction failure of
tailings dams resulting from the near Izu Oshima earthquake, 14 and 15 January
1978. 6th Pan American Conference on Soil Mechanics & Foundation Engineering,
Lima, Bitech Publishing. Vol. 2, pp. 69-80.
Marsal, R.J. (1982) Monitoring of embankment dam behaviour. Proceedings of the 14th
International Congress on Large Dams, Rio de Janiero, ICOLD. pp. 1441-1467
(Q.52 R.84).
Marsal, R.J. and Alberro, J.A. (1976) Performance of dams built in Mexico.
Proceedings of the 12th International Congress on Large Dams, Mexico City,
ICOLD. pp. 791-798 (C.2).
Marsal, R.J. and Ramirez de Arellano, L. (1964) El Infiernillo rockfill dam.
Proceedings of the 8th International Congress on Large Dams, Edinburgh, ICOLD.
Vol. 3 pp. 855-877 (Q.31 R.18).
Marsal, R.J. and Ramirez de Arellano, L. (1967) Performance of El Infiernillo dam
1963-1966. A.S.C.E., Journal of the Soil Mechanics and Foundations Division, Vol.
93 (SM4), pp. 265-298.
References Page R40

Marsal, R.J. and Ramirez de Arellano, L. (1972) Eight years of observations at El


Infiernillo dam. Proceedings, ASCE Speciality Conference on Performance of
Earth and Earth Retaining Structures, Purdue University, Lafayette, ASCE, New
York. Vol. 1 (Part I), pp. 703-722.
Marsal, R.J. and Tamez, G. (1955) Earth dams in Mexico. Design construction and
performance. Proceedings of the 5th International Congress on Large Dams, Paris,
ICOLD. pp. 1123-1178 (C.30).
Marsal, R.L. (1973) Mechanical properties of rockfill. In Embankment Dam
Engineering (Casagrande Volume) (Hirschfeld and Poulos ed.), John Wiley and
Sons, New York, pp. 109-200.
Marsland, A. and Powell, J.J.M. (1977) The behaviour of a trial bank constructed to
failure on soft alluvium of the River Thames. Proceedings, International
Symposium on Soft Clay, (Brenner & Brand ed.) Bangkok, Asian Institute of
Technology. pp. 505-525.
Martin, G.R. (1998) "Post-liquefaction" shear strength from laboratory and field tests.
National Science Foundation Workshop on Shear Strength of Liquefied Soils,
(Stark, Olson, Kramer and Youd ed.) Urbana, Illinois, National Science
Foundation. pp. 40-52.
Marulanda, A. and Pinto, N.L. (2000) Recent experience on design, construction, and
performance of CFRD dams. In Concrete Face Rockfill Dams, J. Barry Cooke
Volume (Mori, Sobrinho, Dijkstra, Guocheng and Borgatti ed.), Beijing, pp. 279-
315.
Materon, B. (1985a) Alto Anchicaya dam - ten years performance. Proceedings of the
Symposium on Concrete Face Rockfill Dams - Design, Construction and
Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE New York. pp.
73-87.
Materon, B. (1985b) Construction of Foz do Areia dam. Proceedings of the Symposium
on Concrete Face Rockfill Dams - Design, Construction and Performance, (Cooke
and Sherard ed.) Detroit, Michigan, ASCE New York. pp. 192-207.
Matsui, I. (1976) Measurement on relative displacement between different materials at
boundary zones and a consideration on a proposed analysis for fill dams taking into
account the measurement. Proceedings of the 12th International Congress on Large
Dams, Mexico City, ICOLD. Vol. 4, pp. 895-917 (C.8).
McConnell, A.D., Paré, J.J., Verma, N.S. and Rattue, A.B. (1982) Materials for
construction methods for the dam and dyke embankments of the LG-4 project.
References Page R41

Proceedings of the 14th International Congress on Large Dams, Rio de Janiero,


ICOLD. pp. 123-144 (Q.55 R.8).
Melbourne and Metropolitan Board of Works (1972) Greenvale Reservoir construction
report.
Melbourne and Metropolitan Board of Works (1975) Sugarloaf reservoir project,
design report. Water Supply Division.
Melbourne and Metropolitan Board of Works (1981) Notes on rock materials used at
Winneke dam. ANCOLD Bulletin, Vol 59, pp. 57-61.
Melbourne and Metropolitan Board of Works (1985) Toorourrong - Yan Yean system
description and data and notes for inspection report. File No. 505/002/0502,
Structural Surveillance Section, Water Supply Division.
Melbourne and Metropolitan Board of Works (1986) Upper Yarra reservoir:
Surveillance report. MMBW, Water Supply Division, Surveillance Group.
Melbourne and Metropolitan Board of Works (1987a) Cardinia reservoir surveillance
report. Report No. 505/008/0502, MMBW, Structural Surveillance Section,
Systems Planning Division.
Melbourne and Metropolitan Board of Works (1987b) Thomson reservoir: Surveillance
report. Report No. 505/012/0502, MMBW, Structural Surveillance Section,
System Planning Division.
Melbourne and Metropolitan Board of Works (1988) Greenvale Reservoir surveillance
report. Water Supply Division.
Melbourne and Metropolitan Board of Works (1995) Sugarloaf reservoir, monitoring
report (7 June 1995). Water Assets Section.
Mesri, G. and Choi, Y.K. (1979) Excess pore water pressures during consolidation.
Proceedings, Sixth Asian Conference on Soil Mechanics and Foundation
Engineering, Singapore, Vol. 1, pp. 151-154.
Middlebrooks, T.A. (1940) Fort Peck slide. A.S.C.E. Transactions (December), pp. 723-
742.
Midgley, D.C. (1979) The failure of a slimes dam at Bafokeng; hydrological aspects and
a barrier to further escape of slimes. The Civil Engineer in South Africa, Vol 21 (6),
pp. 151-154.
Ministry of Works and Development New Zealand (1987) Ohau B Power Project:
Earthworks construction report.
Mitchell, J.K. (1964) Shearing resistance of soils as a rate process. A.S.C.E., Journal of
the Soil Mechanics and Foundations Division, Vol 90 (SM1), pp. 29-61.
References Page R42

Mitchell, J.K. (1986) Practical problems from surprising soil behaviour, 20th Terzaghi
Lecture. A.S.C.E., Journal of the Geotechnical Division, Vol 112 (3), pp. 255-289.
Mitchell, J.K. (1993) Fundamentals of Soil Behavior. John Wiley & Sons, Inc.
Mitchell, R.J. and Eden, W.J. (1972) Measured movements of clay slopes in the Ottawa
area. Canadian Journal of Earth Sciences, Vol 9, pp. 1001-1013.
Mitchell, R.J. and Markell, A.R. (1974) Flowsliding in sensitive soils. Canadian
Geotechnical Journal, Vol 11, pp. 11-31.
Mitchell, R.J. and Williams, D.R. (1981) Induced failure of an instrumented clay slope.
Proceedings of the 10th International Conference on Soil Mechanics and
Foundation Engineering, Stockholm, Balkema, Rotterdam. Vol. 3, pp. 479-484.
Mitchell, W.R., Fidler, J. and Fitzpatrick, M.D. (1968) Rowallan and Parangana rockfill
dams. The Journal of The Institution of Engineers, Australia (Oct-Nov), pp. 239-
251.
Mitchell, W.R. and Fitzpatrick, M.D. (1979) An incident at Rowallan dam. Proceedings
of the 13th International Congress on Large Dams, New Delhi, ICOLD. pp. 195-
210 (Q.49 R.14).
Montanez, L.E., Hacelas, J.E. and Castro, A.J. (1993) Design of Aguamilpa dam.
Proceedings of the International Symposium on of High Earth Rockfill Dams,
Beijing, ICOLD. pp. 337-363.
Moreno, E. and Alberro, J.A. (1982) Behaviour of the Chicoasen dam: construction and
first filling. Proceedings of the 14th International Congress on Large Dams, Rio de
Janiero, ICOLD. pp. 155-182 (Q.52 R.9).
Morfitt, B.J. (1984) Preliminary analysis of static stability of San Luis dam abutments.
Technical Memorandum No. SL-III-222-1, United States Bureau of Reclamation,
Division of Design, Embankment Dams Section.
Morgenstern, N.R. (1994) Report on the Kwun Lung Lau landslide of 23 July 1994.
Volume 1, for Geotechnical Engineering Office, Civil Engineering Department,
Hong Kong Government.
Morgenstern, N.R. (2000) Performance in geotechnical practice. The first Lumb
lecture. pp. 1-58.
Mori, R.T. (1999) Deformation and cracks in concrete face rockfill dams. Proceedings
of the Second Symposium on Concrete Face Rockfill Dams, Florianopolis, Brazil,
Brazilian Committee on Dams. pp. 49-61.
References Page R43

Mori, R.T. and Pinto, N.L.S. (1988) Analysis of deformations in concrete face rockfill
dams to improve face movement prediction. Proceedings of the 16th International
Congress on Large Dams, San Francisco, ICOLD. pp. 27-34 (Q.61 R.2).
Morris, S.B. (1939) Discussion on Galloway paper: The design of rock-fill dams.
Transactions, ASCE, Vol. 104, pp. 35-38.
Morse, A.R. (1995) White Spur dam, Surveillance Report No. 1 (1986 - 1995). Report
ENE-0010-SF-013, Hydro-Electric Commission Tasmania, Civil and Water
Resources Engineering Group.
Morse, A.R. and Ward, M.J. (1989) Anthony power development, White Spur Dam,
Civil Design Report (record). Report CDR 523, Hydro-Electric Commission
Tasmania, Civil Design Section.
Muir Wood, D. (1990) Soil Behaviour and Critical State Soil Mechanics. Cambridge
University Press, Cambridge, pp. 462.
Murley, K.A. and Cummins, P.J. (1982) Design considerations of materials during
construction of Dartmouth dam, Australia. Proceedings of the 14th International
Congress on Large Dams, Rio de Janiero, ICOLD. pp. 627-641 (Q.55 R.37).
Myrvoll, F., Larsen, S., Sande, A. and Romslo, N.B. (1985) Field instrumentation and
performance observations for the Vatnedalsvatn dams. Proceedings of the 15th
International Congress on Large Dams, Lausanne, ICOLD. pp. 1039-1069 (Q.56
R.56).
Nadilo, J.P. (1983) Happy Valley dam. Report on stability of embankment. Ref. No.
83/27, South Australian Government, Engineering and Water Supply Department,
Design Services Branch.
Nagarkar, P.K., Kulkarni, R.P. and Kulkarni, M.V. (1978) Case histories of two dams in
Maharshtra. Proceedings of Geocon Conference on Geotechnical Engineering,
New Delhi, Vol. 1, pp. 64-69.
Nagarkar, P.K., Kulkarni, R.P., Kulkarni, M.V. and Kulkarni, D.G. (1981) Failures of a
monozone earth dam of expansive clay. Proceedings of the 10th International
Conference on Soil Mechanics and Foundation Engineering, Stockholm, Vol. 3,
pp. 491-494.
Nakagawa, K., Komada, H. and Kanazawa, K. (1985) Application of improved analysis
on the pore pressure behaviour within the impervious zone to a prototype rockfill
dam. Proceedings of the 15th International Congress on Large Dams, Lausanne,
ICOLD. Vol. 1, pp. 537-578 (Q.56 R.26).
References Page R44

National Science Foundation (1998) Shear strength of liquefied soils. National Science
Foundation Workshop on Shear Strength of Liquefied Soils, (Stark, Olson, Kramer
and Youd ed.) Urbana, Illinois, National Science Foundation.
Naylor, D.J., Maranha das Neves, E., Mattar, D. and Veiga Pinto, A.A. (1986)
Prediction of construction performance of Beliche dam. Geotechnique, Vol. 36 (3),
pp. 359-376.
Naylor, D.J., Maranha, J.R., Maranha das Neves, E. and Veiga Pinto, A.A. (1997) A
back-analysis of Beliche dam. Geotechnique, Vol. 47 (2), pp. 221-233.
Naylor, D.J., Tong, S.L. and Shahkarami, A.A. (1989) Numerical modelling of
saturation shrinkage. Proc. Numerical models in Geomechanics NUMOG III,
(Pietruszczak & Pande ed.) Amsterdam, Elsevier. pp. 636-648.
Newland, J.R. and Davidson, J.S. (1979) Chaffey dam: Construction report 1975 -
1979. Water Resources Commission, New South Wales.
Nichols, R.S. (1982) Waste dump stability at Fording Coal Limited in B.C. Proceedings
of the Third International Symposium on Stability in Surface Mining, pp. 795-813.
Nilsson, A. and Norstedt, U. (1991) Evaluation of aging processes in two Swedish
dams. Proceedings of the 17th International Congress on Large Dams, Vienna,
ICOLD. pp. 23-47 (Q.65 R.2).
Nobari, E.S. and Duncan, J.M. (1972a) Effect of reservoir filling on stresses and
movements in earth and rockfill dams. Report TE-72-1, University of California,
Department of Civil Engineering.
Nobari, E.S. and Duncan, J.M. (1972b) Movements in dams due to reservoir filling.
Proceedings, ASCE Speciality Conference on Performance of Earth and Earth
Retaining Structures, Purdue University, Indianna, ASEC, New York. Vol. 1 (Part
1), pp. 797-815.
Nonveiller, E. and Anagnosti, P. (1961) Stresses and deformations in cores of rockfill
dams. Proceedings, 5th International Conference on Soil Mechanics and
Foundation Engineering, Paris, Vol. 2, pp. 673-680.
Norem, H., Locat, J. and Schieldrop, B. (1990) An approach to the physics and the
modelling of submarine flowslide. Marine Geotechnology, Vol. 9, pp. 93-111.
Nose, M. (1969) Kisenyama pumped storage project. Water Power, Vol. 21 (11), pp.
411-420.
Nose, M. (1982) Present trends in construction and operation of dams in Japan.
Proceedings of the 14th International Congress on Large Dams, Rio de Janiero,
ICOLD. pp. 693-728 (GP/RS.1).
References Page R45

Nutt, K.D. (1975) Maroon dam embankment and foundation. Proceedings of the 2nd
Australia New Zealand Conference on Geomechanics, Brisbane, IEAust, Sydney.
pp. 129-133.
Oborn, L.E. (1985) Cobb Power Station SEED Examination: Report on Engineering
Geology. for Electricity Corporation of New Zealand Limited (Electricorp).
Oikawa, T., Harita, K. and Mizuno, M. (1997) Sealing works for an old river bed gravel
stratum on a reservoir bank. Proceedings of the 19th International Congress on
Large Dams, Florence, ICOLD. Vol. 3, pp. 449-487 (Q.74 R.28).
Okubo, R., Nakahira, E. and Miyagawa, T. (1988) Automated embankment
management for the construction of rockfill dam. Proceedings of the 16th
International Congress on Large Dams, San Francisco, ICOLD. Vol. 3, pp. 1003-
1025 (C.5).
Olsauskas, A.A., Kotowicz, M. and Paton, R.J. (1967a) Blowering dam and
appurtenant works, Contract No. 20,094. Construction progress report: 30th June
1967. Snowy Mountains Hydro Electric Authority.
Olsauskas, A.A., Lawrence, A.I. and Paton, R.J. (1967b) Blowering dam and
appurtenant works, Contract No. 20,094. Construction progress report: 31st
December 1967. Snowy Mountains Hydro Electric Authority.
Olsauskas, A.A., Lawrence, A.I. and Paton, R.J. (1968) Blowering dam and
appurtenant works, Contract No. 20,094. Construction progress report: Final
Report. Snowy Mountains Hydro Electric Authority.
Olsauskas, A.A., Perkins, R.W., Macoun, K. and Paton, R.J. (1966) Blowering dam and
appurtenant works, Contract No. 20,094. Construction progress report: 31st
December 1966. Snowy Mountains Hydro Electric Authority.
Olsen, R.S. (1997) Cyclic liquefaction based on the cone penetrometer test. Proceedings
of the NCEER Workshop on Evaluation of Liquefaction Resistance of Soils, (Youd
and Idriss ed.) Salt Lake City, Utah, National Centre for Earthquake Engineering
Research. pp. 225-276.
Olson, S.M., Stark, T.D. and Castro, G. (2000) 1907 static liquefaction and flow failure
of the north dike of Wachusett Dam. A.S.C.E., Journal of Geotechnical and
Geoenvironmental Engineering, Vol 126 (12), pp. 1184-1193.
Opus International Consultants Ltd (1998a) Benmore power station: Deformation
Survey No. 22B. Contract No. EC95002 Report No. 236, for Electricity
Corporation of New Zealand, Southern Generation.
References Page R46

Opus International Consultants Ltd (1998b) Pukaki Lake control: Deformation Survey
No. 17B. Contract No. EC95002, Report No. 243, for Electricity Corporation New
Zealand, Southern Generation.
Opus International Consultants Ltd (1999) Ruataniwha dam & spillway: Deformation
Survey No. 26. Contract No. EC95002 Report No. 272, for Electricity Corporation
of New Zealand.
Oziz, U., Basmaci, E. and Harmancioglu, N. (1990) Ataturk nears completion. Water
Power & Dam Construction, September, pp. 12-16.
Pagano, L., Desideri, A. and Vinale, F. (1998) Interpreting settlement profiles of earth
dams. A.S.C.E., Journal of Geotechnical and Geoenvironmental Engineering, Vol.
124 (10), pp. 923-932.
Paré, J.J. (1984) Earth dam behaviour on the La Grande project. Proceedings of the
International Conference on Safety of Dams, (Serafim ed.) Coimbra, Portugal,
Balkema, Boston. pp. 153-160.
Paré, J.J., Boncompain, B., Konrad, J.M. and Verma, N.S. (1982) Embankment
compaction and quality control at James Bay hydroelectric development.
Transportation Research Record 897 pp. 8-15.
Paré, J.J., Verma, N.S., Keira, H.M.S. and McConnell, D. (1984) Stress-deformation
predictions for the LG-4 main dam. Canadian Geotechnical Journal, Vol. 21, pp.
213-222.
Parkin, A.K. (1971) Application of rate analysis to settlement problems involving creep.
Proceedings of the First Australia New Zealand Conference on Geomechanics,
Melbourne, Vol. 1, pp. 138-143.
Parkin, A.K. (1977) The compression of rockfill. Australian Geomechanics Journal,
Vol. G7, pp. 33-39.
Partyka, G. and Bowling, A.J. (1984) P.R.P.D., Lower Pieman dam, Rockfill and filter
material control testing. Report 4413-9, Hydro-Electric Commission Tasmania,
Civil Engineering Laboratories.
Paterson, S.J. (1971) Engineering geology of the Lemonthyme hydro-electric scheme,
Tasmania. I.E. Aust., Civil Engineering Transactions (April), pp. 17-24.
Patrick, J.G. (1967) Post-construction behavior of Round Butte dam. A.S.C.E., Journal
of the Soil Mechanics and Foundations Division, Vol. 93 (SM4), pp. 251-263.
Peng, Z. (2000) Analysis of the deformation of Xibeikou CFRD in eight years of
operation. Proceedings of the International Symposium on Concrete Faced Rockfill
Dams, Beijing, ICOLD. pp. 555-564.
References Page R47

Penman, A.D.M. (1985) The failure of Acu Dam (Technical Note 6). Proceedings of the
Symposium on Failures in Earthworks, ICE, London, Thomas Telford, London. pp.
414-416.
Penman, A.D.M. (1986) On the embankment dam. Geotechnique, Vol 36 (3), pp. 303-
348.
Penman, A.D.M. (1988) Remedial measures at Matahina dam following the Edgecumbe
earthquake of 2 March 1987. for Electricity Corporation of New Zealand
(Electricorp).
Penman, A.D.M., Burland, J.B. and Charles, J.A. (1971) Observed and predicted
deformations in a large embankment dam during construction. Proceedings of the
Institution of Civil Engineers, Vol 49 (May), pp. 1-21.
Peter J. Burgess & Associates Pty Ltd (1991) Khancoban dam safety review. for Snowy
Mountains Hydro-Electric Authority.
Peterson, O.W. (1939) Discussion on Galloway paper: The design of rock-fill dams.
Transactions, ASCE, Vol. 104, pp. 40-41.
Peterson, R., Iverson, N.L. and Rivard, P.J. (1957) Studies of several dam failures on
clay foundations. Proceedings of the Fourth International Conference on Soil
Mechanics and Foundation Engineering, London, Vol. 2, pp. 348-352.
Picarelli, L. (2000) Mechanisms and rates of slope movements in fine grained soils.
Proceedings of the International Conference on Geotechnical and Geological
Engineering (GeoEng2000), Melbourne, Technomic, Lancaster. Vol. 1, pp. 1618-
1670.
Picarelli, L., Urciuoli, G. and Russo, C. (2000) Mechanics of slope deformation and
failure in stiff clays and clay shales as a consequence of pore pressure fluctuation.
Proceedings of the 8th International Symposium on Landslides, Cardiff, Wales,
Balkema.
Pierson, T.C. and Costa, J.E. (1987) A rheologic classification of subaerial sediment-
water flows. Geological Society of America, Reviews in Engineering Geology, Vol
7, pp. 1-12.
Pilot, G. (1972) Study of five embankment failures on soft soils. Proceedings, ASCE
Speciality Conference on Performance of Earth and Earth Retaining Structures,
Purdue University, Lafayette, Vol. 1 (1), pp. 81-100.
Pinkerton, I.L. (1984) Review of Happy Valley dam to assess structural adequacy and
safety. For the Engineering and Water Supply Department of the South Australian
Government.
References Page R48

Pinkerton, I.L. and Paton, R.J. (1968) Design and construction of Geehi dam. The
Journal of the Institution of Engineers, Australia, Vol. 40 (March), pp. 33-48.
Pinto, N.L.d.S., Blinder, S. and Toniatti, N.B. (1993) Foz do Areia and Segredo CFRD
dams - 12 years evolution. Proceedings of the International Symposium on High
Earth-Rockfill Dams, (Jiang, Zhang & Qin ed.) Beijing, Vol. 1, pp. 381-398.
Pinto, N.L.d.S. and Marques Filho, P.L. (1998) Estimating the maximum face slab
deflection in CFRDs. Hydropower & Dams, No. 6.
Pinto, N.L.d.S., Marques Filho, P.L. and Maurer, E. (1985a) Foz do Areia dam - design,
construction and behaviour. Proceedings of the Symposium on Concrete Face
Rockfill Dams - Design, Construction and Performance, (Cooke and Sherard ed.)
Detroit, Michigan, ASCE New York. pp. 173-191.
Pinto, N.L.d.S., Marques Filho, P.L. and Maurer, E. (1985b) Segredo dam - basic
design aspects. Proceedings of the Symposium on Concrete Face Rockfill Dams -
Design, Construction and Performance, (Cooke and Sherard ed.) Detroit,
Michigan, ASCE New York. pp. 587-593.
Pinto, N.L.d.S., Materon, B. and Marques Filho, P.L. (1982) Design and performance of
Foz do Areia concrete membrane as related to basalt properties. Proceedings of the
14th International Congress on Large dams, Rio de Janiero, ICOLD. pp. 873-906
(Q.55 R.51).
Pinto, N.L.S. and Marques Filho, P.L. (1985) Discussion of Clements paper: Post-
construction deformation of rockfill dams. A.S.C.E., Journal of Geotechnical
Engineering, Vol. 111 (12), pp. 1472-1475.
Pitman, T.D., Robertson, P.K. and Sego, D.C. (1994) Influence of fines on the collapse
surface of loose sands. Canadian Geotechnical Journal, Vol 31, pp. 728-739.
Pope, R.J. (1967) Evaluation of Cougar dam embankment performance. A.S.C.E.,
Journal of the Soil Mechanics and Foundations Division, Vol. 93 (SM4), pp. 231-
250.
Pope, R.J. (1969) Closure: Evaluation of Cougar dam embankment performance.
A.S.C.E., Journal of the Soil Mechanics and Foundations Division, Vol. 95 (SM4),
pp. 1116.
Potts, D.M., Dounias, G.T. and Vaughan, P.R. (1990) Finite element analysis of
progressive failure of Carsington embankment. Geotechnique, Vol. 40 (1), pp. 79-
101.
Potts, D.M., Kovacevic, N. and Vaughan, P.R. (1997) Delayed collapse of cut slopes in
stiff clay. Geotechnique, Vol. 47 (5), pp. 953-982.
References Page R49

Poulos, H.G., Booker, J.R. and Ring, G.J. (1972) Simplified calculations of
embankment deformations. Soils and Foundations, Tokyo, Vol 12 (4), pp. 1-17.
Poulos, H.G. and Davis, E.A. (1974) Elastic Solutions for Soil and Rock Mechanics.
Wiley, New York.
Poulos, H.G., Lee, C.Y. and Small, J.C. (1990) Predicted and observed behaviour of a
test embankment on Malaysian soft clays. Research Report R620, The University
of Sydney, NSW, Australia.
Poulos, S.J. (1998) Written statement: Shear strength of liquefied soils from laboratory
and field tests. Discussion. National Science Foundation Workshop on Shear
Strength of Liquefied Soils, (Stark, Olson, Kramer and Youd ed.) Urbana, Illinois,
National Science Foundation. pp. 152-156.
Public Works and Services Department NSW (1998) Mangrove Creek dam - Post
construction embankment behaviour. Report DC98072, DPWS NSW, Dams and
Civil Section.
Public Works Department N.S.W. (1990) Gosford Wyong water supply, Mangrove
Creek dam monitoring survey (unpublished).
Public Works Department NSW (1992a) Burrendong dam design report. PWD NSW,
Dams and Civil Section (for Department of Water Resources).
Public Works Department NSW (1992b) Wyangala dam design report. PWD NSW,
Dams and Civil Section (for Department of Water Resources).
Public Works Department NSW (19XX) Copeton dam design report. PWD NSW,
Dams Branch (for Department of Water Resources).
Pun, W.K. and Yeo, K.C. (1995) Report on the investigation of the 23 July and 7
August 1994 landslides at Milestone 14.5 Castle Peak Road. Advisory Report
ADR 1/95, Geotechnical Engineering Office, Civil Engineering Department, Hong
Kong.
Queensland Department of Natural Resources (1997) Bjelke-Petersen dam: Five yearly
safety inspection, 10th to 12th June 1997. QDNR, State Waters Projects.
Queensland Department of Natural Resources (1998) Peter Faust dam: Five yearly
safety inspection. QDNR, State Waters Projects.
Queensland Department of Primary Industries (1994) Design of Peter Faust dam.
Report No. 69/R13, QDPI, Water Resources, Water Production Division.
Queensland Department of Primary Industries (1995) Split-Yard Creek dam - Data
book.
References Page R50

Queensland Department of Primary Industries (1996) Split-Yard Creek dam safety


review. QDPI, Water Resources (report for Austa Electric).
Queensland Irrigation and Water Supply Commission (1976) The geology of Wivenhoe
dam site: Part 1.
Queensland Water Resources Commission (1979a) A report on the construction of
Wivenhoe dam Stage II. Draft Report.
Queensland Water Resources Commission (1979b) Schedule of rate contract for third
stage construction of Wivenhoe dam. Contract No. 2120 (Volumes II & III),
QWRC, Design and Construction.
Queensland Water Resources Commission (1986a) Barker-Barambah Irrigation
Project. Schedule of rates contract. Construction of Bjelke-Petersen dam and
appurtenant works. Contract No. 2580 (Volumes II & IV).
Queensland Water Resources Commission (1986b) Sources of construction materials
for construction of Bjelke-Petersen dam.
Queensland Water Resources Commission (1987) Post-construction report on Split-
Yard Creek dam (Wivenhoe dam and pumped storage hydro-electric project). Ref.
No. CD/1, QWRC, Construction Division.
Queensland Water Resources Commission (19XXa) Maroon Dam Safety Evaluation
Data Book.
Queensland Water Resources Commission (19XXb) Split-Yard Creek dam - Draft
design report.
Quinlan, P. (1993) King River power development, Crotty Dam, Performance Report.
Report CDR 610, Hydro-Electric Commission Tasmania, Consulting Business
Unit.
Ramalho-Ortigao, J.A., Lacerda, W.A. and Werneck, M.L.G. (1983a) The behaviour of
the instrumentation of an embankment on soft clay. Proceedings, International
Symposium on Field Measurements in Geomechanics, (Kovari ed.) Zurich, Vol. 1,
pp. 703-717.
Ramalho-Ortigao, J.A., Werneck, M.L.G. and Lacerda, W.A. (1983b) Embankment
failure on clay near Rio de Janiero. A.S.C.E., Journal of Geotechnical Engineering,
Vol 109 (11), pp. 1460-1479.
Ramirez de Arellano, L. and Gomez, E.M. (1972) Field measurements at La Angostura
cofferdams. Proceedings, ASCE Speciality Conference on Performance of Earth
and Earth Retaining Structures, Purdue University, Indianna, ASCE, New York.
Vol. 1 (Part 1), pp. 779-796.
References Page R51

Rao, K.L. (1957) Behaviour of recent earth dams and levees in India. Proceedings of the
4th International Conference on Soil Mechanics and Foundation Engineering,
London, Butterworths Scientific Publications, London. Vol. 2, pp. 361-367.
Rao, K.L. and Wadhwa, H.L. (1958) Observation at Hirakud earth dam. Proceedings of
the 6th International Congress on Large Dams, London, ICOLD. pp. 471-482
(Q.21 R.53).
Read, S.A.L. (1976) Upper Waitaka power development scheme. Pukaki Lake control.
Engineering geological completion report. Report No. EG268, New Zealand
Geological Survey, Dept Scientific and Industrial Research, Engineering Geology
Section.
Ready, O.T. (1910) Construction of the Belle Fourche dam. Engineering Record, Vol.
61 (14), pp. 466-469.
Regalado, G., Materon, B., Ortega, J.W. and Vargas, J. (1982) Alto Anchicaya concrete
face rockfill dam behaviour of the concrete face membrane. Proceedings of the
14th International Congress on Large Dams, Rio de Janiero, ICOLD. Vol. 4, pp.
517-535 (Q.55 R.30).
Regan, P.J. (1997) Performance of concrete-faced rockfill dams of the Pacific Gas &
Electric Company. Non-Soil Water Barriers for Embankment Dams, Proceedings of
the 17th Annual USCOLD Lecture Series, San Diego, California, USCOLD. pp.
149-162.
Regan, W.M. (1980) Winneke Reservoir, main dam geological report on construction.
Engineering Geology Section Report 81/489, Melbourne and Metropolitan Board of
Works.
Reinhold, J.J. (1969) The geology of the western area: Greenvale Reservoir site.
Report No. 3, Melbourne and Metropolitan Board of Works, Civil Engineering
Laboratories.
Resendiz, D. and Romo, M.P. (1972) Analysis of embankment deformations.
Proceedings, ASCE Speciality Conference on Performance of Earth and Earth
Retaining Structures, Purdue University, Indianna, ASCE, New York. Vol. 1 (Part
I), pp. 817-836.
Riemer, M.F. (1998) Written statement: Shear strength of liquefied soils from
laboratory and field tests. Discussion. National Science Foundation Workshop on
Shear Strength of Liquefied Soils, (Stark, Olson, Kramer and Youd ed.) Urbana,
Illinois, National Science Foundation. pp. 232-237.
References Page R52

Riemer, W. (2001) Personnel communication with Dr. W. Riemer (Geological


Consultant).
Robertson, P.K., Woeller, D.J. and Finn, W.D.L. (1992) Seismic cone penetration test
for evaluating liquefaction potential under cyclic loading. Canadian Geotechnical
Journal, Vol. 29 (3), pp. 686-695.
Robertson, P.K. and Wride, C.E. (1997) Cyclic liquefaction and its evaluation based on
the SPT and CPT. Proceedings of the NCEER Workshop on Evaluation of
Liquefaction Resistance of Soils, (Youd and Idriss ed.) Salt Lake City, Utah,
National Centre for Earthquake Engineering Research. pp. 41-87.
Robinson, P. (1979) Murchison dam - Control testing of rockfill placing. Report 4352-
1, Hydro-Electric Commission Tasmania, Civil Engineering Laboratories.
Rocke, G. (1993) Technical Note: Investigation of the failure of Carsington Dam.
Geotechnique, Vol. 43 (1), pp. 175-180.
Rodine, J.D. (1974) Analysis of the mobilisation of debris flows. Ph.D. dissertation,
Stanford University, California.
Rogers, B.T., Been, K., Hardy, M.D., Johnson, G.J. and Hachey, J.E. (1990) Re-
analysis of Nerlerk B-67 berm failures. Proceedings of the Canadian Geotechnical
Engineering Conference, Laval University, Quebec, Vol. I pp. 227-237.
Rogers, R.L. and Pearce, K.G. (1991) Effect of the foundation on design, construction
and filling of Split-Yard Creek dam. Proceedings of the 17th International
Congress on Large Dams, Vienna, ICOLD. pp. 1577-1599 (Q.66 R.84).
Rowe, P.W. (1991) A reassessment of the causes of the Carsington embankment failure.
Geotechnique, 41 (3), pp. 395-421.
Rowe, R.K., Gnanendran, C.T., Landva, A.O. and Valsangkar, A.J. (1995) Construction
and performance of a full-scale geotextile reinforced test embankment, Sackville,
New Brunswick. Canadian Geotechnical Journal, Vol 32, pp. 514-534.
Rowe, R.K., Gnanendran, C.T., Landva, A.O. and Valsangkar, A.J. (1996) Calculated
and observed behaviour of a reinforced embankment over soft compressible soil.
Canadian Geotechnical Journal, Vol 33, pp. 324-338.
Rudd, R.T. (1979) The failure of a slimes dam at Bafokeng; the Bafokeng disaster and
its legal implications. The Civil Engineer in South Africa, Vol 21 (6), pp. 146-150.
Saboya, F., Barbosa, R. and Vasconcelos, A. (2000) The influence of the left abutment
geometry on the behaviour of Xingo rockfill dam. Proceedings of the International
Symposium on Concrete Faced Rockfill Dams, Beijing, ICOLD. pp. 565-572.
References Page R53

Saboya, F.J. and Byrne, P.M. (1993) Parameters for stress and deformation analysis of
rockfill dams. Canadian Geotechnical Journal, Vol. 30, pp. 690-701.
Saito, M. (1965) Forecasting the time of occurrence of a slope failure. Proceedings of
the 6th International Conference on Soil Mechanics and Foundation Engineering,
Montreal, Vol. 2, pp. 537-541.
Sakamoto, T., Takebayashi, S., Nakamura, A. and Yasuda, N. (1994) Safety assessment
based on the observed behaviour of zoned rockfill dams. Proceedings of the 18th
International Congress on Large Dams, Durban, ICOLD. pp. 925-953 (Q.68 R.55).
Sassa, K. (1988) Special lecture: Geotechnical model for the motion of landslides.
Proceedings of the Fifth International Symposium on Landslides, (Bonnard ed.)
Lausanne, Switzerland, Balkema. Vol. 1, pp. 37-55.
Savage, J.L. (1931) Report on repairs to Belle Fourche dam, Belle Fourche Project.
United States Department of the Interior, Bureau of Reclamation.
Sbeghen, B. (1990) Geological investigations for Maroon dam safety evaluation. Ref.
No. PP2351, Queensland Water Resources, Planning Division, Engineering
Geology Section.
Scheidegger, A.E. (1973) On the prediction of the reach and velocity of catastrophic
landslides. Rock Mechanics, Vol 5, pp. 231-236.
Schiffman, R.L., Pane, V. and Gibson, R.E. (1984) The theory of one-dimensional
consolidation of saturated clays, IV: An overview on non-linear finite strain
sedimentation and consolidation. Proceedings of the ASCE Symposium on
Sedimentation Consolidation Models, Prediction and Validation, (Yong &
Townsend ed.) San Francisco, ASCE, New York. pp. 1-29.
Schmidt, L.A.J. (1958) Rockfill dams: Performance and maintenance of Dix River dam.
A.S.C.E., Journal of the Power Division, Vol. 84 (PO3), pp. 1683-1 to 29.
Schober, W. (1967) Behaviour of the Gepatsch rockfill dam. Proceedings of the 9th
International Congress on Large Dams, Istanbul, ICOLD. pp. 677-699 (Q.34
R.39).
Schober, W. (1970) Behaviour interior stress distribution of the Gepatsch rockfill dam.
Proceedings of the 10th International Congress on Large Dams, Montreal, ICOLD.
Vol. 1, pp. 169-187 (Q.36 R.10).
Schuyler, J.D. (1912) Reservoirs for Irrigation, Water-Power and Domestic Water
Supply. John Wiley and Sons, New York.
References Page R54

Schwab, H.H. (1979) The Gepatsch rockfill dam - Analysis relating to long-term
behaviour (1962 to 1978) (in German). Osterreichische Wasserwirtschaft, Vol. 31
(5/6), pp. 202-210.
Seed, H.B. (1979) Soil liquefaction and cyclic mobility evaluation for level ground
during earthquakes. A.S.C.E., Journal of the Geotechnical Engineering Division,
Vol 105 (GT2), pp. 210-255.
Seed, H.B. (1987) Design problems in soil liquefaction. A.S.C.E., Journal of
Geotechnical Engineering, Vol 113 (8), pp. 827-845.
Seed, H.B., Idriss, I.M. and Arango, I. (1983) Evaluation of liquefaction potential using
field performance data. A.S.C.E., Journal of Geotechnical Engineering, Vol 109
(3), pp. 458-482.
Seed, H.B., Lee, K.L. and Idriss, I.M. (1969) Analysis of Sheffield Dam failure.
A.S.C.E., Journal of the Soil Mechanics and Foundations Division, Vol 95 (SM6),
pp. 1453-1490.
Seed, H.B., Lee, K.L., Idriss, I.M. and Makdisi, F.I. (1975) The slides in the San
Fernando dams during the earthquake of February 9, 1971. A.S.C.E., Journal of the
Geotechnical Engineering Division, Vol 101 (GT7), pp. 651-688.
Seed, H.B., Tokimatsu, K., Harder, L.F. and Chung, R. (1985) Influence of SPT
procedures in soil liquefaction resistance evaluations. A.S.C.E., Journal of
Geotechnical Engineering, Vol 115 (12), pp. 1425-1445.
Seed, R.B. and Harder, L.F. (1990) SPT-based analysis of cyclic pore pressure
generation and undrained residual strength. Proceedings, H. Bolton Seed Memorial
Symposium, (Duncan ed.) BiTech Publishers. Vol. 2, pp. 351-376.
Shaw, E.D. (1953) Construction of the Eildon project. Unpublished paper (?).
Sherard, J.L. (1953) Influence of soil properties and construction methods on the
performance of homogeneous earth dams. Ph.D. Thesis, Harvard University.
Sherard, J.L. (1973) Embankment dam cracking. In Embankment Dam Engineering
(Casagrande Volume) (Hirschfeld and Poulos ed.), John Wiley and Sons, New
York, pp. 271-353.
Sherard, J.L. and Cooke, J.B. (1987) Concrete-face rockfill dam: I. Assessment. Journal
of Geotechnical Engineering, ASCE, Vol 113 (10), pp. 1096-1112.
Sherard, J.L., Woodward, R.J., Gizienski, S.F. and Clevenger, W.A. (1963) Earth and
Earth-Rock Dams. John Wiley and Sons, New York, pp. 725.
Shiraiwa, K. and Takahashi, Y. (1985) Construction of Tokachi dam on the Tokachi
River. Civil Engineering in Japan, Vol. 24, pp. 133-148.
References Page R55

Siddle, H.J., Wright, M.D. and Hutchinson, J.N. (1996) Rapid failures of colliery spoil
heaps in the South Wales coalfield. Quarterly Journal of Engineering Geology, Vol
29, pp. 103-132.
Sierra, J.M., Ramirez, C.A. and Hacelas, J.E. (1985) Design features of Salvajina dam.
Proceedings of the Symposium on Concrete Face Rockfill Dams - Design,
Construction and Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE
New York. pp. 266-285.
Silvas, F. and de Groot, M.B. (1995) Flow slides in the Netherlands: experience and
engineering practice. Canadian Geotechnical Journal, Vol 32, pp. 1086-1092.
Silveira, J.F. and Sardinha, A.E. (1999) Behaviour of Ita CFRD at the end of the
construction period. Proceedings of the Second Symposium on Concrete Face
Rockfill Dams, Florianopolis, Brazil, Brazilian Committee on Dams. pp. 37-48.
Sinclair Knight Merz (1995) Kangaroo Creek dam - Report on 1st stage of safety
evaluation. For Engineering and Water Supply Department of South Australia.
Singh, A. and Mitchell, J.K. (1968) General stress-strain-time function for soil. ASCE,
Journal of the Soil Mechanics and Foundations Division, Vol 94 (SM1), pp. 21-46.
Sinha, D.S. (1968) Dams in distress. Irrigation and Power, Vol. 25 (3), pp. 301-320.
Sivathayalan, S. and Vaid, Y.P. (1998) Truly undrained response of granular soils with
no membrane-penetration effects. Canadian Geotechnical Journal, Vol 35, pp.
730-739.
Skempton, A.W. (1948) The rate of softening of stiff, fissured clays. Proceedings of the
Second International Conference on Soil Mechanics and Foundation Engineering,
Rotterdam, Vol. 2, pp. 50-53.
Skempton, A.W. (1954) The pore pressure coefficients A and B. Geotechnique, Vol. 4
(4), pp. 143-147.
Skempton, A.W. (1964) Long-term stability of clay slopes. Geotechnique, Vol. 14 (2),
pp. 77-101.
Skempton, A.W. (1977) Slope stability of cuttings in brown London Clay. Proceedings
of the Ninth International Conference on Soil Mechanics and Foundation
Engineering, Tokyo, Vol. 3, pp. 261-270.
Skempton, A.W. (1985) Residual strength of clays in landslides, folded strata and the
laboratory. Geotechnique, Vol. 35 (1), pp. 3-18.
Skempton, A.W. (1986) Standard penetration test procedures and the effects in sands of
overburden pressure, relative density, particle size, aging and overconsolidation.
Geotechnique, Vol. 36 (3), pp. 425-447.
References Page R56

Skempton, A.W. (1990) Historical development of British embankment dams to 1960.


Clay Barriers for Embankment Dams, London, Thomas Telford, London. pp. 15-
52.
Skempton, A.W. and La Rochelle, P. (1965) The Bradwell slip: A short-term failure in
London clay. Geotechnique, Vol. 15 (3), pp. 221-242.
Skempton, A.W., Schuster, R.L. and Petley, D.J. (1969) Joints and fissures in the
London clay at Wraysbury and Edgeware. Geotechnique, Vol. 19 (2), pp. 205-217.
Skempton, A.W. and Vaughan, P.R. (1993) The failure of Carsington Dam.
Geotechnique, Vol. 43 (1), pp. 151-173.
Sladen, J.A., D'Hollander, R.D. and Krahn, J. (1985a) The liquefaction of sands, a
collapse surface approach. Canadian Geotechnical Journal, Vol 22, pp. 564-578.
Sladen, J.A., D'Hollander, R.D., Krahn, J. and Mitchell, D.E. (1985b) Back analysis of
the Nerlerk berm liquefaction slides. Canadian Geotechnical Journal, Vol 22, pp.
579-588.
Sladen, J.A. and Hewitt, K.J. (1989) Influence of placement method on the in situ
density of hydraulic sand fills. Canadian Geotechnical Journal, Vol 26, pp. 453-
466.
Smith, D. and Hungr, O. (1992) Failure behaviour of large rockslides. Report No. 16-
11-6, Thurber Engineering Ltd., Report to the Geological Survey of Canada and BC
Hydro and Power Authority.
Snowy Mountains Engineering Corporation (1971) Cardinia Creek dam and
appurtenant works: Design report. for Melbourne and Metropolitan Board of
Works.
Snowy Mountains Engineering Corporation (1975) Dartmouth dam project: Interim
design report. for State Rivers and Water Supply Commission, Victoria.
Snowy Mountains Engineering Corporation (1986) Tooma dam: Report on safety
review. for Snowy Mountains Hydro-Electric Authority.
Snowy Mountains Engineering Corporation (1990) Island Bend, Tantangara, Guthega
and Eucumbene dams: Review of instrumentation results and records. for Snowy
Mountains Hydro-Electric Authority.
Snowy Mountains Engineering Corporation (1995) Hope Valley dam safety evaluation
report. Document No. 35501.001, for Engineering and Water Supply Department
of South Australia.
References Page R57

Snowy Mountains Engineering Corporation (1996a) Lake Nillahcootie: Safety


Surveillance 1995 and formal Dam Safety Inspection 18 April 1996. Document No.
44519.200, for Goulburn-Murray Water.
Snowy Mountains Engineering Corporation (1996b) Lake William Hovell: Report on
Safety Surveillance and Behaviour 1995. Document No. 44519.130, for Goulburn
Murray Water.
Snowy Mountains Engineering Corporation (1996c) Tullaroop Reservoir: Report on
Safety Surveillance and Behaviour 1995. Document No. 44519.300, for Goulburn-
Murray Water.
Snowy Mountains Engineering Corporation (1996d) Yan Yean dam: Safety review of
dam and outlets. Document No. 34509.01, for Melbourne Water Corporation.
Snowy Mountains Engineering Corporation (1997a) Lake Dartmouth: Report on safety
surveillance and behaviour 1996. for Goulburn-Murray Water.
Snowy Mountains Engineering Corporation (1997b) Yan Yean dam stability review:
further investigation and analysis. Document No. 44615.400, for Melbourne Water
Corporation.
Snowy Mountains Engineering Corporation (1998a) Lake Bellfield: Report on Safety
Surveillance and Behaviour 1997 and formal Dam Safety Inspection 26 February
1998. Document No. 44812.100, for Wimmera Mallee Water.
Snowy Mountains Engineering Corporation (1998b) Upper Yarra dam embankment
safety review. Docume nt No. 44818.200, for Melbourne Water Corporation.
Snowy Mountains Engineering Corporation (1998c) Yan Yean dam: Embankment
rehabilitation design report. Document No. 44823.200, for Melbourne Water
Corporation.
Snowy Mountains Engineering Corporation (1998d) Lake Eppalock Geotechnical
Investigation - Phase 2. Document No. 44821.110 CDB11JK8, for Goulburn
Murray Water.
Snowy Mountains Engineering Corporation (1999a) Eildon design review - stage 1.
Document No. 44930, for Goulburn-Murray Water.
Snowy Mountains Engineering Corporation (1999b) Hope Valley dam study. Document
No. 35901.001, for South Australian Water Corporation.
Snowy Mountains Engineering Corporation (2002) Happy Valley dam: Interim report
on the numerical modelling. Document No. 35001.002.02, for South Australian
Water Corporation.
References Page R58

Snowy Mountains Engineering Corporation (unpublished) Thomson dam: Draft design


report.
Snowy Mountains Hydro-Electric Authority (1964) Schedule of rates contract for
Blowering dam and appurtenant works: Contract No. 20,094.
Snowy Mountains Hydro-Electric Authority (1998a) Eucumbene dam and saddle dam:
1998 annual report. SMHEA, Operations Dam Safety Section.
Snowy Mountains Hydro-Electric Authority (1998b) Khancoban dam 1998 annual
report. SMHEA, Operations Dam Safety Section.
Snowy Mountains Hydro-Electric Authority (1998c) Talbingo dam 1998 annual report.
SMHEA, Operations Dam Safety Section.
Sobrinho, J.A., Sadinha, A.E., Albertoni, S.C. and Dijkstra, H.H. (2000) Development
aspects of CFRD in Brazil. In Concrete Face Rockfill Dams, J. Barry Cooke
Volume (Mori, Sobrinho, Dijkstra, Guocheng and Borgatti ed.), Beijing, pp. 153-
175.
Sobrinho, J.A., Sardinha, A.E. and Fernandes, A.M. (1999) Ita dam - design and
construction. Proceedings of the Second Symposium on Concrete Face Rockfill
Dams, Florianopolis, Brazil, Brazilian Committee on Dams. pp. 377-390.
Soga, K. and Mitchell, J.K. (1996) Rate dependent deformation of structured natural
clays. ASCE Geotechnical Special Publication No. 61, Measuring and Modeling
Time Dependent Soil Behavior, (Sheahan & Kaliakin ed.) Washington, D.C.,
ASCE, New York. pp. 243-257.
Sokolov, I.B., Marchuk, A.N., Kuznetsov, V.S., Aleksandrovskaya, E.A., Kuzmin,
K.K., Pavlov, V.L., Tsaryov, A.I. and Alipov, V.V. (1985) Analysis and
interpretation of measurement data illustrated by the construction and staged
commissioning of the Sayano-Shushenskaya and Nurek hydro power plants.
Proceedings of the 15th International Congress on Large Dams, Lausanne,
ICOLD. pp. 1471-1482 (Q.56 R.75).
Sonu, J. (1985) Performance of instrument in rockfill dams in Korea. Proceedings of the
15th International Congress on Large Dams, Lausanne, ICOLD. pp. 855-868 (Q.56
R.46).
South Australian Water Corporation (1995) Dam Safety Committee inspection of
Kangaroo Creek dam.
Souza, R.J.B., Cavalcanti, A.J.C.T., Silva, S.A. and Silveira, J.F. (1999) Xingo concrete
face rockfill dam behaviour of the dam on the left abutment. Proceedings of the
References Page R59

Second Symposium on Concrete Face Rockfill Dams, Florianopolis, Brazil,


Brazilian Committee on Dams. pp. 143-157.
Sowers, G.F., Davie, J., Soenarno and Mansoer, M.N. (1993) Jatiluhur Dam: problems
and rehabilitation. Proc. of the Conference on Geotechnical Practice in Dam
Rehabilitation, (Anderson ed.) ASCE, New York. pp. 17-34.
Sowers, G.F., Williams, R.C. and Wallace, T.S. (1965) Compressibility of broken rock
and the settlement of rockfills. Proceedings of the 6th International Conference on
Soil Mechanics and Foundation Engineering, Montreal, University of Toronto
Press. Vol. 2, pp. 561-565.
Soydemir, C. and Kjœrnsli, B. (1975) A treatise on the performance of rockfill dams
with unyielding foundations in relation to the design of Storvass dam. Report No.
53203, Norwegian Geotechnical Institute.
Soydemir, C. and Kjœrnsli, B. (1979) Deformations of membrane-faced rockfill dams.
Proceedings, 7th European Conference on Soil Mechanics and Foundation
Engineering, Brighton, England, Vol. 3, pp. 281-284.
Speedie, M.G. (1948) Investigations and designs for Eildon Dam enlargement. Journal
of the Institution of Engineers, Australia, Vol. 20 (7-8), pp. 81-91.
Squier, L.R. (1968) Discussion: Evaluation of Cougar dam embankment performance
(by Pope 1967). A.S.C.E., Journal of the Soil Mechanics and Foundations Division,
Vol. 94 (SM3), pp. 780-783.
Squier, L.R. (1970) Load transfer in earth and rockfill dams. A.S.C.E., Journal of the
Soil Mechanics and Foundations Division, Vol. 96 (SM1), pp. 213-233.
Stafford, C.T. and Weatherburn, D.C. (1958) Glenbawn dam: construction. The Journal
of The Institution of Engineers, Australia, Vol. 30 (12), pp. 333-351.
Stark, T.D. and Duncan, J.M. (1987) Mechanisms of strength loss in stiff clays.
Virginia Polytechnic Institute and State University, Dept. of Civil Engineering,
Geotechnical Engineering.
Stark, T.D. and Duncan, J.M. (1991) Mechanisms of strength loss in stiff clays.
A.S.C.E., Journal of Geotechnical Engineering, Vol. 117 (1), pp. 139-155.
Stark, T.D. and Mesri, G. (1992) Undrained shear strength of liquefied sands for
stability analysis. A.S.C.E., Journal of Geotechnical Engineering, Vol 118, (11), pp.
1727-1747.
State Rivers and Water Supply Commission Victoria (1983) Cairn Curran Dam:
Preliminary design review report. Major Projects Design Division.
References Page R60

Stateler, J.N. (1983) Structural behaviour report: Bradbury dam, Cachuma Project,
California (Mid-Pacific Region). United States Department of the Interior Bureau
of Reclamation, Division of Dam Safety, Structural Behaviour Branch.
Steele, I.C. and Cooke, J.B. (1958) Rockfill dams: Salt Springs and Lower Bear River
concrete face dams. A.S.C.E., Journal of the Power Division, Vol. 84 (PO4), pp.
1737-1 to 43.
Steele, I.C. and Dreyer, W. (1939) Discussion on Galloway paper: The design of rock-
fill dams. Transactions, ASCE, Vol. 104, pp. 53-63.
Stille, H., Fredriksson, A. and Broms, B.B. (1976) Analysis of a test embankment
considering the anisotropy of the soil. Proceedings, Second International
Conference on Numerical Methods in Geomechanics, (Desai ed.) Virginia, ASCE,
New York. Vol. 2, pp. 611-622.
Stoutjesdijk, T.P., de Groot, M.B. and Lindenberg, J. (1998) Flow slide prediction
method: influence of slope geometry. Canadian Geotechnical Journal, Vol 35, pp.
43-54.
Straw, A.J., Bennett, C.A., Reid, G.B., Druce, M.J. and Chenhall, G.A. (1985)
Windemere dam main embankment: Construction report 1980 - 1985. Department
of Water Resources, New South Wales.
Stroman, W.R., Beene, R.R.W. and Hull, A.M. (1984) Clay shale foundation slide at
Waco dam, Texas. Proceedings of the International Conference on Case Histories
in Geotechnical Engineering, (Prakash ed.) Rolla, USA, Vol. 2, pp. 579-586.
Stroman, W.R. and Karbs, H.E. (1985) Monitoring and analyses of pore pressures; clay
shale foundation, Waco Dam, Texas. 15th International Congress on Large Dams,
Lausanne, ICOLD. pp. 599-620 (Q.56 R.29).
Su, W.X. and Miller, H.D.S. (1995) Waste pile stability and debris flow formation.
Rock Mechanics, (Daemen & Schultz ed.) Balkema. pp. 831-836.
Sun, H.W. (1999) Review of fill slope failures in Hong Kong. GEO Report 96,
Geotechnical Engineering Office, Civil Engineering Department, Government of
Hong Kong.
Svenson, D., Burgess, J.B. and Hosking, A.D. (1964) Report on sources of construction
materials for Blowering dam and appurtenant works. Snowy Mountains Hydro
Electric Authority.
Tait, G.A. (1963) Some construction aspects of the Benmore earth dam. Proceedings of
the 4th Australia-New Zealand Conference on Soil Mechanics and Foundation
Engineering, Adelaide, IEAust, Sydney. pp. 76-80.
References Page R61

Takahashi, M. and Nakayama, K. (1973) The effect of regional conditions in Japan on


design and construction of impervious elements of rockfill dams. Proceedings of
the 11th International Congress on Large Dams, Madrid, ICOLD. pp. 501-524
(Q.42 R.29).
Tavenas, F. (1984) Landslides in Canadian sensitive clays - A state-of-the-art.
Proceedings of the Fourth International Symposium on Landslides, Toronto, Vol. 1,
pp. 141-153.
Tavenas, F., Changnon, J.Y. and La Rochelle, P. (1971) The Saint-Jean-Vianney
landslide: Observations and eyewitnesses accounts. Canadian Geotechnical
Journal, Vol 8, pp. 463-478.
Tavenas, F. and Leroueil, S. (1977) Effects of stresses and time on yielding of clays.
Proceedings, Ninth International Conference on Soil Mechanics and Foundation
Engineering, Tokyo, Vol. 1, pp. 319-326.
Tavenas, F. and Leroueil, S. (1980) The behaviour of embankments on clay
foundations. Canadian Geotechnical Journal, Vol 17, pp. 236-260.
Tavenas, F. and Leroueil, S. (1981a) Creep and failure of slopes in clay. Canadian
Geotechnical Journal, Vol 18, pp. 106-120.
Tavenas, F. and Leroueil, S. (1981b) Reply: The behaviour of embankments on clay
foundations. Canadian Geotechnical Journal, Vol 18, pp. 462-466.
Tavenas, F., Leroueil, S., La Rochelle, P. and Roy, M. (1978) Creep behaviour of an
undisturbed lightly overconsolidated clay. Canadian Geotechnical Journal, Vol 15,
pp. 402-423.
Tavenas, F., Mieussens, C. and Bourges, F. (1979) Lateral displacements in clay
foundations under embankments. Canadian Geotechnical Journal, Vol 16, pp. 532-
550.
Taylor, R.K. (1984) Composition and engineering properties of British colliery discard.
National Coal Board, Mining Department.
Tedd, P., Charles, J.A. and Holton, I.R. (1997a) Settlement of old embankment dams: a
guide to measurement and interpretation. Dams and Reservoirs, Vol 7 (March), pp.
18-23.
Tedd, P., Charles, J.A., Holton, I.R. and Robertshaw, A.C. (1994) Deformation of
embankment dams due to changes in reservoir level. Proceedings of the 13th
International Conference on Soil Mechanics and Foundation Engineering, New
Delhi, Vol. 3, pp. 951-954.
References Page R62

Tedd, P., Charles, J.A., Holton, I.R. and Robertshaw, A.C. (1997b) The effect of
reservoir drawdown and long-term consolidation on the deformation of old
embankment dams. Geotechnique, Vol 47 (1), pp. 33-48.
Tedd, P., Claydon, J.R. and Charles, J.A. (1990) Deformation of Ramsden dam during
reservoir drawdown and refilling. Proceedings of the 6th Conference of the British
Dams Society (BNCOLD). The Embankment Dam., Nottingham, Thomas Telford,
London. pp. 171-176.
Tedd, P. and Holton, I.R. (1987) Ramsden dam: interim report on the investigation of
the downstream fill and crest settlements. Note N28/87, Building Research
Establishment.
Tedd, P., Robertshaw, A.C. and Holton, I.R. (1993) Investigation of an old embankment
dam with an upstream clay blanket and central clay core. Dams and Reservoirs, Vol
3 (1), pp. 6-9.
Terzaghi, K. (1950) Mechanism of landslides. Engineering Geology (Berkley) Volume,
The Geological Society of America (November), pp. 83-123.
Terzaghi, K. (1956) Varieties of submarine slope failures. Proceedings of the Eighth
Texas Conference on Soil Mechanics and Foundation Engineering, University of
Texas, Austin, pp. 1-41.
Terzaghi, K. (1958) Design and performance of the Sasumua dam. Proceedings of The
Institution of Civil Engineers, Vol. 9 (April), pp. 369-394.
Terzaghi, K. (1960) Discussion on 1958 paper by Steele & Cooke: Rockfill dams: Salt
Springs and Lower Bear River concrete face dams. ASCE, Transactions, Vol. 125
(Part II), pp. 139-148.
Terzaghi, K. and Peck, R.B. (1948) Soil mechanics in engineering practice. John Wiley
and Sons, New York.
Torrence, J.K. (1996) On the development of high sensitivity: Mineralogical
requirements and constraints. Proc. 7th International Symposium on Landslides,
(Senneset ed.) Trondheim, Norway, Balkema, Rotterdam. Vol. 1, pp. 491-496.
Trak, B. and Lacasse, S. (1996) Soils susceptible to flow slides and associated failure
mechanisms. Proceedings of the Seventh International Symposium on Landslides,
(Senneset ed.) Trondheim, Norway, Balkema, Rotterdam. Vol. 1, pp. 497-506.
Treiber, I.F. (1958a) Compaction methods adopted for the construction of Rosshaupten
dam, their effectiveness, and the behaviour of the impervious loam core.
Proceedings of the 6th International Congress on Large Dams, London, ICOLD.
pp. 123-137 (Q.22 R.8).
References Page R63

Treiber, I.F. (1958b) Measurement and observations on Rosshaupten dam. Proceedings


of the 6th International Congress on Large Dams, London, ICOLD. pp. 215-231
(Q.21 R.5).
Troncoso, J.H. (1988) Evaluation of seismic behavior of hydraulic fill structures.
Hydraulic Fill Structures, ASCE Speciality Conference, Geotechnical Special
Publication No. 21, (Van Zyl and Vick ed.) Colorado, ASCE, New York. pp. 475-
491.
Troncoso, J.H. (2000) Failure mechanisms and shear failures of tailings dams under
earthquake loadings. Proceedings of the International Conference on Geotechnical
and Geological Engineering (GeoEng2000), Melbourne, Vol. 1, pp. 1254-1262.
Troncoso, J.H., Vergara, A. and Avendano, A. (1993) The seismic failure of Barahona
tailings dam. Proc. Third International Conference on Case Histories in
Geotechnical Engineering, St. Louis, Missouri, Vol. 3, pp. 1473-1479.
United States Department of the Interior Bureau of Reclamation (1946) Schedules,
specifications and drawings. Horsetooth Reservoir, Colorado-Big Thompson
Project, Colorado. Specification No. 1275.
United States Department of the Interior Bureau of Reclamation (1951) Enders dam:
Final embankment report.
United States Department of the Interior Bureau of Reclamation (1953) Adaminaby
dam: Report on design of earth embankment. Earth Dams Section, Dams Branch,
Design and Construction Division (for Snowy Mountains Hydro-Electric Authority
and Department of Public Works New South Wales).
United States Department of the Interior Bureau of Reclamation (1959) Cachuma dam:
Technical record of design and construction.
United States Department of the Interior Bureau of Reclamation (1963a) Final
construction report: Navajo dam. Colorado River Storage Project. Navajo Unit
Project Office.
United States Department of the Interior Bureau of Reclamation (1963b) Steinaker dam:
Technical record of design and construction.
United States Department of the Interior Bureau of Reclamation (1965) Technical
record of design and construction. Trinity River division features of the Central
Valley Project, California. Volumes I - Design.
United States Department of the Interior Bureau of Reclamation (1966a) Navajo dam
and reservoir: Technical record of design and construction.
References Page R64

United States Department of the Interior Bureau of Reclamation (1966b) Technical


record of design and construction. Trinity River division features of the Central
Valley Project, California. Volumes II - Construction.
United States Department of the Interior Bureau of Reclamation (1972) Design
considerations for construction of Pueblo dam, Fryingpan-Arkansas Project,
Colorado.
United States Department of the Interior Bureau of Reclamation (1974) San Luis dam:
Technical record of design and construction.
United States Department of the Interior Bureau of Reclamation (1995) Performance
parameters for San Justo dam. Technical Memorandum No. JU-8311-1.
United States Department of the Interior Bureau of Reclamation (1996) Performance
parameters for Belle Fourche dam. Technical Memorandum No. BF-8311-1.
United States Department of the Interior Bureau of Reclamation (1997) Performance
parameters for the Horsetooth Reservoir dams. Technical Memorandum No. JT-
8312-1.
United States Department of the Interior Bureau of Reclamation (1998) Performance
parameters for Meeks Cabin dam (draft report). Technical Memorandum (Draft)
No. MS-8312-2.
United States Department of the Interior Bureau of Reclamation (19xx) Enders dam:
Technical record of design and construction.
United States Geological Survey (190X) Advertisement, proposal and specifications.
Belle Fourche project, South Dakota. Dam and distribution canals, third contract.
Specification No. 56, Department of the Interior, Reclamation Service.
USCOLD (1975) Lessons from Dam Incidents USA. United States Committee on Large
Dams, Committee on Failures and Accidents to Large Dams, ASCE, New York, pp.
387.
USCOLD (1988) Lessons from Dam Incidents USA-II. United States Committee on
Large Dams, Subcommittee on Dam Incidents and Accidents of the Committee on
Dam Safety, ASCE, New York, pp. 222.
Uthayakumar, M. and Vaid, Y.P. (1998) Static liquefaction of sands under multiaxial
loading. Canadian Geotechnical Journal, Vol 35, pp. 273-283.
Vaid, Y.P. and Campanella, R.G. (1977) Time dependent behavior of undisturbed clay.
A.S.C.E., Journal of the Geotechnical Engineering Division, Vol 103 (GT7), pp.
693-709.
References Page R65

Vaid, Y.P. and Eliadorani, A. (1998) Instability and liquefaction of granular soils under
undrained and partially drained tests. Canadian Geotechnical Journal, Vol 35, pp.
1053-1062.
Vail, A.J. (1972) A Report on the Po Shan Road landslide. The Commission of Enquiry
into the Rainstorm Disasters of June 1972, The Public Works Department,
Government of Hong Kong.
Van Gassen, W. and Cruden, D.M. (1989) Momentum transfer and friction in the debris
of rock avalanches. Canadian Geotechnical Journal, Vol 26, pp. 623-628.
Vanicek, I. (1982) Simple non-standard laboratory tests before and during construction
of Dalesice dam. Proceedings of the 14th International Congress on Large Dams,
Rio de Janiero, ICOLD. Vol. 1, pp. 605-610 (Q.52 R.38).
Varnes, D.J. (1978) Slope movement types and processes. In Landslides: analysis and
control, Special Report No. 176, Transportation Research Board. (Schuster &
Krizek ed.), National Academy of Sciences, Washington D.C., pp. 11-33.
Varnes, D.J. (1982) Time-deformation relations in creep to failure of earth materials.
Proceedings of the Seventh Southeast Asian Geotechnical Conference, (McFeat-
Smith and Lumb ed.) Hong Kong, Vol. 2, pp. 107-130.
Vasconcelos, A.A. and Eigenheer, L.P. (1985) The Xingo rockfill dam. Proceedings of
the Symposium on Concrete Face Rockfill Dams - Design, Construction and
Performance, (Cooke and Sherard ed.) Detroit, Michigan, ASCE New York. pp.
559-565.
Vaunet, J., Leroueil, S. and Faure, R. (1994) Slope movements: a geotechnical
perspective. Proceedings of the 7th International Congress of the International
Association of Engineering Geology, Lisbon, pp. 1637-1646.
Verdugo, R., Castillo, P. and Briceno, L. (1995) Initial soil structure and steady state
strength. Proceedings of the Conference on Earthquake Geotechnical Engineering,
(Ishihara ed.) Balkema. pp. 209-214.
Verma, N.S., Paré, J.J., Boncompain, B., Garneau, R. and Rattue, A.B. (1985)
Behaviour of the LG-4 main dam. Proceedings of the 11th International
Conference on Soil Mechanics and Foundation Engineering, San Francisco,
Balkema, Brookfield. Vol. 4, pp. 2049-2054.
Villegas, F. (1982) Difficulties during construction of the Punchina cofferdam.
Fourteenth International Congress on Large Dams, Rio de Janeiro, ICOLD. pp.
1081-1101 (Q.55 R.60).
References Page R66

Villegas, F. (1984) Incipient failures of dams built with tropical residual soils and their
behavior during seismic events. Proceedings of the International Conference on
Safety of Dams, Coimbra, pp. 39-43.
Volk, P.L. (1987) Glenbawn dam enlargement: Geological construction report. Water
Conservation and Irrigation Commission NSW.
Von Thun, J.L. (1988) San Luis Dam upstream slide. Proceedings of the 11th
International Conference on Soil Mechanics and Foundation Engineering, San
Francisco, Balkema, Brookfield. Vol. 5, pp. 2593-2598.
Von Thun, L. (1982) Back analysis of the San Luis slide with additional attention to the
Qs surface and the deformed section. Draft Report, Technical Memorandum No.
SL-II-222-2, United States Bureau of Reclamation, Division of Design,
Embankment Dams Section.
W.A. Wahler & Associates (1973) Analysis of coal refuse dam failure, Middle Fork
Buffalo Creek, Saunders, West Virginia. USBM Contract S0122084, for U.S. Dept.
of the Interior, Bureau of Mines.
Wagener, F.M., Craig, H.J., Blight, G., McPhail, G., Williams, A.A.B. and Strydom,
J.H. (1998) The Merriespruit tailings dam failure - A review. Proc. Tailings and
Mine Waste '98, Fort Collins, pp. 925-952.
Walker, F.C. and Harber, W.G. (1961) Design of the Trinity dam, an earthfill structure
537 feet high. Proceedings of the 5th International Conference on Soil Mechanics
and Foundation Engineering, Paris, Vol. 2, pp. 743-748.
Walker, W.L. and Duncan, J.M. (1984) Lateral bulging of earth dams. A.S.C.E., Journal
of Geotechnical Engineering, Vol. 110 (7), pp. 923-937.
Wang, P., Shen, Y. and Wang, Y. (1993) Design and construction of Xibeikou concrete-
faced rockfill dam. Proceedings of the International Symposium on High Earth-
Rockfill Dams, (Jiang, Zhang & Qin ed.) Beijing, Vol. 1, pp. 418-429.
Wang, Y., Chen, J. and Shen, Y. (1988) Xibeikou concrete face rockfill dam.
Proceedings of the 16th International Congress on Large Dams, San Francisco,
ICOLD. pp. 1075-1090 (Q.61 R.57).
Watakeekul, S., Roberts, G.J. and Coles, A.J. (1985) Khao Laem - a concrete face
rockfill dam on karst. Proceedings of the Symposium on Concrete Face Rockfill
Dams - Design, Construction and Performance, (Cooke and Sherard ed.) Detroit,
Michigan, ASCE New York. pp. 336-361.
Water Conservation and Irrigation Commission NSW (1964) Burrendong dam: Report
on main wall construction.
References Page R67

Water Conservation and Irrigation Commission NSW (1982) Glenbawn dam


enlargement. Core material for embankments: Existing embankment material.
Water Resources Commission NSW (1978) Copeton dam: Surveillance report.
Water Resources Commission NSW (1979) Chaffey dam: Construction control, Design
summary to 4-1-79.
Water Resources Commission NSW (1992) Burrendong dam surveillance report.
Report No. TS No. 93.055.
Watson, J.D. (1956) Earth movement affecting L.T.E. railway in deep cutting east of
Uxbridge. Proceedings of the Institution of Civil Engineers, Vol. 5, Part 2 (27
March), pp. 302-331.
West, R.P. (1962) Waco Dam slide: Its cause and correction. Engineering News Record
(2 August), pp. 34-36.
Wieczorek, G.F. (1987) Effect of rainfall intensity and duration on debris flows in
central Santa Cruz mountains, California. Reviews in Engineering Geology,
Geological Society of America, Vol. VII, pp. 93-104.
Wieczorek, G.F., Harp, E.L. and Mark, R.K. (1988) Chapter 8. Debris flows and other
landslides in San Mateo, Santa Cruz, Contra Costa, Alameda, Napa, Solano,
Sonoma Lake and Yolo Counties, and other factors influencing debris-flow
distribution. In Landslides, floods, and marine effects of the storm of January 3-5,
1982, in the San Francisco Bay region, California. U.S. Geological Survey
Professional Paper 1434 (Ellen & Wieczorek ed.), U.S. Geological Survey, pp. 163-
183.
Wilkes, P.F. (1972) An induced failure at a trial embankment at King's Lynn, Norfolk,
England. Proceedings, ASCE Speciality Conference on Performance of Earth and
Earth Retaining Structures, Purdue University, Lafayette, Vol. 1 (1), pp. 29-63.
Wilkins, J.K., Mitchell, W.R., Fitzpatrick, M.D. and Liggins, T. (1973) The design of
Cethana concrete face rockfill dam. Proceedings of the 11th International Congress
on Large Dams, Rio de Janiero, ICOLD. pp. 25-43 (Q.42 R.3).
Wilson, N.A. and Scott, H.S. (1957) The design of Glenbawn dam. The Journal of The
Institution of Engineers, Australia, Vol. 29 (Dec), pp. 333-343.
Wilson, S.D. (1973) Deformation of earth and rockfill dams. In Embankment Dam
Engineering (Casagrande Volume) (Hirschfeld and Poulos ed.), John Wiley and
Sons, New York, pp. 365-417.
Wilson, S.D. and Squier, R. (1969) Earth and rockfill dams. Proceedings, 7th
International Conference on Soil Mechanics and Foundation Engineering, Mexico,
References Page R68

Sociodad Mexicana de Mecanica de Suelos, A.C. State of the Art Volume pp. 137-
223.
Wolfskill, L.A. and Lambe, T.W. (1967) Slide in Siburua Dam. Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 93 (SM4), pp. 107-133.
Wong, H.N., Chen, Y.M. and Lam, K.C. (1996) Factual report on the November 1993
natural terrain landslides in three study areas on Lantau Island. Special Project
Report SPR 10/96 (Volumes 1, 2 and 3), Geotechnical Engineering Office, Civil
Engineering Department, Hong Kong.
Wong, H.N. and Ho, K.K.S. (1996) Travel distance of landslide debris. Proceedings of
the Seventh International Symposium on Landslides, (Senneset ed.) Trondheim,
Norway, Balkema, Rotterdam. Vol. 1, pp. 417-422.
Wong, H.N., Ho, K.K.S. and Chan, Y.C. (1997) Assessment of the consequences of
landslides. Landslide Risk Assessment, (Cruden & Fell ed.) Honolulu, Hawaii,
Balkema. pp. 111-149.
Wong, K.L., Kleiner, D.E., Wood, A.M., Geary, M.C. and Oechsel, R.G. (1992) Design
and performance of Bath County Upper Dam and reservoir slopes. Proceeding of
the Conference on Stability and Performance of Slopes and Embankment - II.
Geotechnical Special Publication No. 31, (Seed & Boulanger ed.) Berkeley,
California, ASCE, New York. Vol. 1, pp. 371-386.
Wood, J.J. (1960) Waipapa Power Project: Earth dam construction report. New
Zealand Ministry of Works.
Woodland, A.W. (1969) Geological report on the Aberfan tip disaster of October 21st,
1966. In A selection of technical reports submitted to the Aberfan Tribunal
H.M.S.O., London, Welsh Office, pp. 119-145, (Item 4).
Woodward Clyde (1999) Lake Eppalock main embankment remedial works, design
report. for Goulburn Murray Water.
Works Consultancy Services Ltd (1996a) Benmore power station: Deformation Survey
No. 22, April 1996. WCS Report No. 138, for Electricity Corporation New
Zealand, Southern Generation.
Works Consultancy Services Ltd (1996b) Pukaki Lake control: Deformation Survey No.
17. for Electricity Corporation New Zealand, Southern Generation.
Works Geothermal Ltd (1994) Report on the Type B deformation survey of Matahina
dam. for Electricity Corporation of New Zealand, report prepared for the Waikato
Hydro Group.
References Page R69

Wright, E.L. (1987) Final construction geology report for Pueblo dam. United States
Department of the Interior Bureau of Reclamation, Missouri Basin Regional Office.
Wroth, C.P. and Simpson, B. (1972) An induced failure at a trial embankment. Part II
finite element computations. Proceedings, ASCE Speciality Conference on
Performance of Earth and Earth Retaining Structures, Purdue University,
Lafayette, Vol. 1 (1), pp. 65-79.
Wu, G., Freitas, M.S., Araya, J.A.M. and Huang, Z.Y. (2000a) Planning and
construction of Tianshengqiao 1 CFRD (China). Proceedings of the International
Symposium on Concrete Faced Rockfill Dams, Beijing, ICOLD. pp. 481-496.
Wu, G., Freitas, M.S., Araya, J.A.M., Huang, Z.Y. and Mori, R.T. (2000b)
Tianshengqiao-1 CFRD - monitoring and performance - lessons and new trends for
future CFRDs (China). Proceedings of the International Symposium on Concrete
Faced Rockfill Dams, Beijing, ICOLD. pp. 573-586.
Wu, H., Wujie, Wang, S., Wu, Q. and Cao, K. (2000c) Ten years surveillance of
Chengbing concrete face rockfill dam. Proceedings of the International Symposium
on Concrete Faced Rockfill Dams, Beijing, ICOLD. pp. 595-605.
Wu, Q. and Cao, K. (1993) Concrete faced rockfill dam of Chengbing project.
Proceedings of the International Symposium on High Earth-Rockfill Dams, (Jiang,
Zhang & Qin ed.) Beijing, Vol. 1, pp. 440-447.
Yamamuro, J.A. and Lade, P.V. (1997) Static liquefaction of very loose sands.
Canadian Geotechnical Journal, Vol 34, pp. 905-917.
Yamamuro, J.A. and Lade, P.V. (1998) Steady state concepts and static liquefaction of
silty sands. A.S.C.E., Journal of Geotechnical and Geoenvironmental Engineering,
Vol 124 (9), pp. 868-877.
Yamazumi, A., Hara, N. and Harita, K. (1991) Foundation treatment for Agigawa
rockfill dam constructed on fissured rock. Proceedings of the 17th International
Congress on Large Dams, Vienna, ICOLD. Vol. 3, pp. 831-852 (Q.66 R.47).
Yang, S. (1993) Design of TSQ-1 concrete face rockfill dam. Proceedings of the
International Symposium on High Earth-Rockfill Dams, (Jiang, Zhang & Qin ed.)
Beijing, Vol. 1, pp. 448-543.
Yasunaka, M., Tanaka, T. and Nakano, R. (1985) The behaviour of Fukada earthfill
dam during construction and impounding of the reservoir. Proceedings of the 15th
International Congress on Large Dams, Lausanne, ICOLD. pp. 499-518 (Q.56
R.24).
References Page R70

Yoshimi, Y., Tokimatsu, K. and Hosaka, Y. (1989) Evaluation of liquefaction resistance


of clean sands based on high-quality undisturbed samples. Soils and Foundations,
Vol. 29 (1), pp. 868-877.
Yoshimi, Y., Tokimatsu, K. and Ohara, J. (1994) In situ liquefaction resistance of clean
sands over a wide density range. Geotechnique, Vol. 44 (3), pp. 479-494.
Youd, T.L., Idriss, I.M., Arango, R.D., Castro, G., Christian, J.T., Dobry, R., Finn,
W.D.L., Harder, L.F., Hynes, M.E., Ishihara, K., Koester, J.P., Liao, S.S.C.,
Marcuson, W.F., Martin, G.R., Mitchell, J.K., Moriwaki, Y., Power, M.S.,
Robertson, P.K., Seed, R.B. and Stokoe, K.H. (1997) Summary Report.
Proceedings of the NCEER Workshop on Evaluation of Liquefaction Resistance of
Soils, (Youd and Idriss ed.) Salt Lake City, Utah, National Centre for Earthquake
Engineering Research. pp. 1-40.
APPENDIX A

Terminology and Definitions


for Landslides
Appendix A Page i

TABLE OF CONTENTS

1.0 TERMS AND DEFINITIONS FOR LANDSLIDES ................. A1


1.1 Velocity Classification of Landslides............................................................. A1
1.2 Landslide Classification ................................................................................. A1
1.3 Definition of Symbols Used for Slope and Travel Geometry........................ A5
1.4 Slope Failure Geometry.................................................................................. A6
1.5 Degree of Confinement ................................................................................ A12
Appendix A: Terminology and Definitions Page A1

1.0 TERMS AND DEFINITIONS FOR LANDSLIDES

1.1 VELOCITY CLASSIFICATION OF LANDSLIDES


The post failure velocity of the slide mass is classified according to the system proposed
by the International Union of Geological Sciences (IUGS 1995), as presented in Table
A1.1. Two broad descriptors to define the post failure velocity of landslides are used
throughout the thesis:
i) “Slow Landslide”. The classification of “slow” landslides relates to those
landslides of very slow, slow and moderate post failure velocities as defined by
IUGS (1995), and having an upper velocity limit of 1.8 m/hour.
ii) “Rapid Landslide”. The classification of “rapid” landslides relates to those
landslides of rapid, very rapid and extremely rapid post failure velocity as defined
by IUGS (1995), having a lower velocity limit of 1.8 m/hour. Most of what are
termed “rapid” landslides would classify as very rapid or extremely rapid, having
post failure velocities greater than 3 metres/minute and generally in the order of
metres/sec.

Table A1.1: IUGS (1995) velocity classifications for landslides.


Velocity Description of Velocity limits Velocity in
Classification Velocity mm/sec
7 Extremely rapid > 5 m/sec > 5x10 3
6 Very rapid 3 m/min to 5 m/sec 50 to 5x103
5 Rapid 1.8 m/hour to 3 m/min 0.5 to 50
4 Moderate 13 m/month to 1.8 m/hour 5x10-3 to 0.5
3 Slow 1.6 m/year to 13 m/month 50x10-6 to 5x10-3
2 Very slow 16 mm/year to 1.6 m/year 0.5x10-6 to 50x10-6
1 Extremely slow ≤ 16 mm/year ≤ 0.5x10-6

1.2 LANDSLIDE CLASSIFICATION


The method of classification of a landslide is according to Hutchinson (1988).
However, for landslides where a large portion of the slide mass evacuates the source
Appendix A: Terminology and Definitions Page A2

area post failure, a dual system of classification is used to describe the initiating
landslide in the source area (termed initial slide classification), and to describe the
subsequent movement of the slide in the travel region (termed travel classification).
The main terms used to define the initiating landslide and post failure travel of the slide
are shown in Figure A1.1.

Figure A1.1: Landslide classification system and main slide types.

For classification of the initial landslide, the main classification term (or terms) is
indicative of the failure mechanics of the landslide (either in soils that are dilative or
contractive on shearing) and is generally followed by a description of the surface of
rupture within the general categories of rotational, translational and compound. Dilative
or contractive describes the initial tendency of the soil on the surface of rupture to
increase (dilate) or decrease (contract) in volume under drained shear, or to develop
negative or positive pore water pressures in undrained shear when in a saturated (or near
saturated) condition. The main classification terms for the initial slide classification and
there definitions are:
• Flow slide is used to describe landslides in saturated (or near saturated) soils that are
contractive on shearing, where the failure or rapid acceleration of the slide mass
occurs as a result of a large loss in undrained strength due to static liquefaction on
shearing.
• “Slide of debris” for description of landslides in colluvium, talus and other slope
mantling debris.
Appendix A: Terminology and Definitions Page A3

• Slide through the soil mass is used to describe failures in dilative soils where the
surface of rupture is located through the soil mass.
• Defect Controlled Slide. The term defect controlled slide is used as an initiating
slide classification to describe landslides that are dominantly controlled by defects
in the soil or weathered rock mass.

For the subsequent travel of the landslide, the “travel classification” can be difficult
due to the complex and potentially changing behaviour of the moving slide mass. The
main purpose of the travel classification is to describe the mode of movement of the
slide mass, whether it is dominantly by sliding along a basal surface, in which case the
material carried remains relatively intact, or the movement is more as a turbulent flow.
The travel classification terms used are as follows:
• Flow slide – where the main volume of the slide mass is bodily carried on a
liquefied basal zone. The flow slide travel classification is only applicable to
landslides initially classified as flow slides.
• Debris flow – travel classification to describe turbulent post failure slide movements
of a combination of water, air and debris. The slide mass is a broken up mass of
material that no longer retains its original structure or fabric. The resistance to flow
of a debris flow is a combination of friction (both sliding and internal) and viscous
forces. The term “debris” as used here is a generic term. Several terms are used
that come under the general debris flow classification, they are:
− Debris flow - where used on its own infers that the moving mass is near
saturated. In most cases the pressure of the pore fluid in the debris mass is
greater than hydrostatic because of partial liquefaction due to remoulding and
breakdown during the flow.
− Dry debris flow - where the flowing mass is generally in a dry condition.
− Hyper-concentrated stream-flow is a term used by Pierson and Costa (1987) to
describe a flowing mixture of water and sediment that possesses measurable
yield strength but still appears to flow like a liquid.
• Debris slide – used as a travel classification where the movement occurs essentially
as sliding on a defined basal surface and the slide mass remains relatively intact
(still retains its structure and fabric) during travel. Debris Slide is generally used as
a descriptor for slides that travel beyond the source area, and intact slide for slides
where the post failure deformation is largely confined to the source area.
Appendix A: Terminology and Definitions Page A4

For a number of slides the mode of movement will change with distance from the
source area. Initially a translational slide in a weathered rock mass may occur as basal
sliding on a defect and the slide mass may initially move as a virtually intact body. As
the moving mass is displaced beyond the toe it is likely to break up to some degree. It
may totally disintegrate into a mass of soil and rock pieces and flow down the slope
until it comes to rest. In addition, part of the slide mass may remain within the failure
scarp in a virtually intact condition and part may have exited the failure bowl and totally
disintegrated as it flowed down-slope. Thus the travel stage is complex and difficult to
classify as one particular type of movement.
Given the complexity associated with the travel of some landslides, the predominant
mechanism of movement has been used to classify the travel stage in most cases.
However, in some cases a dual classification system has been used; i.e. debris slide /
debris flow or flow slide / confined debris flow.
Several of the terms used in the adopted system are slightly different to their
interpretation by Hutchinson (1988) and this has been because of the difficulty of using
a dual classification system within the Hutchinson classification system. The terms
debris slide and debris flow cause particular difficulties with their interpretation.
Hutchinson (1988) defines debris slides as predominantly translational “slides of
debris” mantling a slope where the slide material is typically of low cohesion and
undergoes significant break-up during travel. He then differentiates debris flows from
debris slides as flow type movements of essentially wet debris. Johnson and Rodine
(1984) encompass the term debris slide under the general heading of debris flow.
Pierson and Costa (1987) propose a rheologic classification system for sediment-water
flows (Figure A1.2) that would appear to incorporate both debris slides and debris flows
as described above under the “flow” classification system. From the Hutchinson
description of debris slide the Pierson and Costa system would describe them as “grain
flows” under the sub-group of granular flows (flow movements of material in an
essentially dry or with limited water condition).
Appendix A: Terminology and Definitions Page A5

Figure A1.2 : Rheologic classification system for sediment-water flows (Pierson and
Costa 1987)

1.3 DEFINITION OF SYMBOLS USED FOR SLOPE AND TRAVEL


GEOMETRY
The definitions of the terms used to describe the geometry of the landslide (Figure
A1.3) are:
• Slope height, Hs – overall height of the slope, mainly used in relation to cut and fill
slopes;
Appendix A: Terminology and Definitions Page A6

• Slide height, H – height from crest of back-scarp to distal toe of the post-slide debris
deposition;
• Travel distance, L – horizontal distance from crest of back-scarp to distal toe of the
post-slide debris deposition;
• Travel distance angle, a - the angle between the crest of the back-scarp and the
distal toe of the post-slide debris deposition;
• Ltoe – horizontal distance of travel beyond the toe of the initial failure to the distal
toe of the post-slide debris deposition;
• Failure depth, D – vertical depth of the initial failure in the source area, measured
from the ground surface to the surface of rupture. It is generally representative of
close to the maximum depth to the surface of rupture;
• Sources area slope angle, a 1 - slope angle in the slide source area. For cut slopes it
is the cut slope angle (a cut ) in the region of the initial slide, and for fill slopes it is
the fill slope angle (a fill) in the region of the initial slide;
• a 2 - down-slope angle immediately down-slope of the failure zone;
• a 3 - down-slope angle over the distal portion of the travel path. For landslides of
Type 2 slope failure geometry and some Type 3, a 3 is undefined where the
downslope angle is relatively constant for the full travel length beyond the source
area;
• a 4 - applicable to failures in fill slopes only. It is the slope of the natural ground
underlying the fill as defined in Figure A1.4;
• a base – the inclination angle of the surface of rupture. It is generally measured in the
central to toe region of the slide, as shown in Figure A1.5.

The terms cut slope angle (a cut) and fill slope angle (a fill) are used to avoid
ambiguity between a 1 and a 2 for failures in cut (and fill) slopes where the source area of
the failure is either within part of the slope or encompasses the full slope.

1.4 SLOPE F AILURE G EOMETRY


The slope failure geometry is a classification of the overall longitudinal geometry of the
landslide. The classification includes an assessment of the location of the initial failure
within the overall slope and the longitudinal shape of the down-slope travel path
(mainly for slides where the post failure travel distance is largely beyond the source area
Appendix A: Terminology and Definitions Page A7

of the initial failure). Five slope failure geometry classifications (Type 1 to Type 5)
have been used and are defined in Figure A1.4 to Figure A1.8 respectively.

Figure A1.3: Symbol definitions of slope geometry

Figure A1.4: Type 1 slope geometry - initial failure above the toe of a cut or fill slope.
For “rapid” landslides the deposition is predominantly on relatively flat slope (less than
about 5 degrees) at the toe of the slope.
Appendix A: Terminology and Definitions Page A8

Figure A1.5: Type 2 slope geometry - initial failure coincident with toe of slope (e.g. cut
slope) and large decrease in slope angle from a 1 to a 2. Mostly for cut and fill slopes for
which a 2 is near horizontal.

Figure A1.6: Type 3 slope geometry. Initial failure and post failure travel on relatively
uniform or gradually decreasing slope angles.

Figure A1.7: Type 4 slope geometry - post failure travel on changing slope angles from
steep to shallow.
Appendix A: Terminology and Definitions Page A9

Figure A1.8: Type 5 slope failure geometry. The surface of rupture extends below and
daylights beyond the toe of the slope; (a) retained cut or fill slope or (b) deep seated
failure through the foundation of a fill embankment.

The purpose in differentiating between the slope failure geometry is because it can (and
does) have a significant effect on the post failure travel distance and velocity of the
landslide. For example, for Type 1 failures in steep embankment or cut slopes, the toe
of the slide mass can travel relatively large distances at high velocities down the cut or
fill slope.
For landslides where the post failure travel is largely beyond the source area of the
initial failure, slope failure geometries of Types 1 to 4 are appropriate. These cases are
generally landslides of “rapid” post failure velocity. The slope failure geometry
classifications relate to:
• Type 1 – for failures in cut and fill slopes where the slide source area is above the
toe of slope, and the slide mass travels down the cut or fill slope depositing
predominantly on relatively flat slopes (less than about 5 degrees) at the toe of the
slope.
• Type 2 – for failures in cut or fill slopes where the toe of the slide source is
coincident with toe of slope and the slide mass is predominantly deposited on
relatively flat slopes at the toe of the slope (i.e. a 2 is near horizontal).
• Type 3 – for failures where the source area and travel of the slide mass are on
relatively uniform or gradually decreasing slope angles. Type 3 geometries are
Appendix A: Terminology and Definitions Page A10

generally slides in natural slopes, but the term is also used for slides in cut of fill
slopes where the slide mass is deposited on the cut or fill slope.
• Type 4 – for failures where the travel path of the slide mass is characterised by sharp
changes in slope angle and does not classify as Type 1.

Slope Failure Geometries of Type 1, 2 and 4 are characterised by sharp changes in


slope angle that are likely to absorb a significant amount of the energy of the “rapid”
landslide. The slide volume will have an affect on the classification adopted. For
example, small volume failures in road cuttings would classify as either Type 1 or Type
2; however, a larger volume may classify as Type 4 if the travel extends beyond the fill
edge of road; and a very large volume may classify as Type 3 where the scale of the
road cutting is very small in relation to the travel path.
For landslides where the post failure travel is of limited distance beyond the initial
failure geometry and most of the slide mass is retained in the source area (i.e.
predominantly “slow” and intact slides), slope failure geometries of Type 1, 2 and 5 are
appropriate. The slope failure geometry classifications for these slides relate to:
• Type 1 – where the toe of the landslide is located above the toe of the cut or fill
slope, defined in Figure A1.9. Typical examples of Type 1 failures would include:
− Cut slope failures where the failure is located within the weaker upper portion of
the slope or the basal plane of the failure is along a defect that outcrops in the
face of the cut slope;
− Failures in embankment dams with rockfill toe zones, where the failure is
located above the rockfill zone;
− Failures in embankment dams where the basal plane of the failure is coincident
with a weak layer in the embankment, for example a wet layer with low
undrained strength or a more plastic layer;
• Type 2 – where the toe of the landslide is coincident with the toe of the slope (Figure
A1.10). In the case of benched cut or fill slopes this would also include failures
where the toe of the landslide is coincident with the toe of the bench. Case studies
where the toe of the landslide may be only slightly beyond the toe of the slope have
been included in the Type 2 category, and would include failures in embankment
dams or other fills where the failure incorporates the upper portion of the foundation
as shown in Figure A1.10b.
Appendix A: Terminology and Definitions Page A11

• Type 5 – where the toe of the landslide extends below and beyond the toe of the
slope (Figure A1.8). Typical examples of Type 5 failure geometries would include:
− Failures in retained cut slopes where the failure extends below the base of the
retaining wall (Figure A1.8a); and
− Deep seated failures in embankment dams that extend well into the foundation
and daylight well beyond the toe of the slope (Figure A1.8b).

Figure A1.9: Type 1 failure slope geometries. The initial failure is located above the toe
of the slope.

Figure A1.10: Type 2 failure slope geometries. The initial failure is coincident with the
toe of the slope.
Appendix A: Terminology and Definitions Page A12

1.5 DEGREE OF CONFINEMENT


For classification of the degree of confinement of the travel path of the slide mass, two
general categories have been considered; confined and unconfined travel paths (Figure
A1.11). For several groups of slides the term partly confined travel path has also been
used. The definitions for each type of travel path (after Golder Assoc. 1995) are as
follows:
• Confined – the travel path is constrained by the relatively steep side-slopes of a
gully or small valley;
• Partly confined – the travel path is not sharply defined by a topographic depression.
This includes; movement of the slide mass down relatively broad (relative to the
slide mass volume) topographic depressions, movements down relatively small
topographic depressions where only a portion the slide mass is confined but the
depression controls the direction of movement to a large extent, or events where
only a minor (but potentially significant) portion of travel path is confined; and
• Unconfined – the travel path is on open slopes.

Figure A1.11: Typical plan views of the slide travel for (a) confined travel paths and (b)
unconfined or open slope travel paths (Golder Assoc. 1995).
APPENDIX B

“Rapid” Landslides in Soil


Slopes
Appendix B Page i

TABLE OF CONTENTS

1.0 CASE STUDIES OF LANDSLIDES IN SOIL SLOPES OF


“RAPID” POST FAILURE VELOCITY................................... B1
1.1 “Rapid” Landslides from Failures in Cut, Fill and Natural Slopes, Hong
Kong ............................................................................................................... B1
1.2 “Rapid” Landslides from Failures in Granular Fills ...................................... B5
1.3 Strongly Retrogressive Landslides................................................................. B7
1.4 “Rapid” Landslides from Failures in Hydraulic Fill Embankment Dams ...... B7

2.0 MECHANICS OF SHEARING OF CONTRACTIVE


GRANULAR SOILS ................................................................. B27
2.1 Introduction .................................................................................................. B27
2.2 Empirical Field Methods for Evaluation of Liquefaction and Flow
Liquefaction Potential .................................................................................. B27
2.3 Problems with Laboratory Test Methods For Evaluation of Flow
Liquefaction Potential .................................................................................. B32
2.4 Characteristics of Contractive and Dilative Soils Sheared Under
Monotonic Load Conditions......................................................................... B35
2.5 Residual Undrained Shear Strength of Flow Liquefied Soils ...................... B50
Appendix B Page ii

LIST OF TABLES
Table B1.1: Shear strength of mineral infillings commonly found in weathered
rocks (Irfan and Woods 1998; Irfan 1997).............................................................. B1
Table B1.2: Shear strength on relict discontinuities in weathered rocks in Hong
Kong (Irfan and Woods 1998; Irfan 1997).............................................................. B2
Table B1.3: Landslide case studies from Hong Kong.................................................... B9
Table B1.4: Landslide case studies from Hong Kong (Ayotte and Hungr 1998) ........ B13
Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles ............ B14
Table B1.6: Flow slides from tailings impoundments – cases initiated by static
liquefaction. ........................................................................................................... B20
Table B1.7: Flow slides from tailings impoundments – cases initiated by dynamic
liquefaction. ........................................................................................................... B22
Table B1.8: Case studies of flow slides in sensitive clays ........................................... B23
Table B1.9: “Rapid” landslides in submarine slopes (Edgars and Karlsrud 1983)...... B24
Table B1.10: Failure case studies of hydraulic fill embankment dams, sub-
aqueous fills and Wachusett Dam. ........................................................................ B25
Appendix B Page B1

1.0 CASE STUDIES OF LANDSLIDES IN SOIL SLOPES OF


“RAPID” POST FAILURE VELOCITY

1.1 “RAPID ” LANDSLIDES FROM FAILURES IN CUT, FILL AND NATURAL SLOPES ,
HONG KONG
Table B1.1 and Table B1.2 present the shear strength properties of mineral infillings
and along relict discontinuities in weathered rocks in Hong Kong published by Irfan
(1997) and Irfan and Woods (1998).
The case study data for failures in cut, fill and natural slopes from Hong Kong that
developed into landslides of “rapid” post failure velocity are presented in Table B1.3
and Table B1.4.

Table B1.1: Shear strength of mineral infillings commonly found in weathered rocks
(Irfan and Woods 1998; Irfan 1997)
Mineral and Nature Strength Parameters
Kaolinite – remoulded φ′r = 12 to 22 degrees
Kaolinite φ′sat = 0.4 to 0.6 φdry
Kaolinite – intact, precut surface φ′r = 12 degrees
Illite – remoulded φ′r = 6.5 to 11 degrees
Montmorillonite φ′r = 4 to 11 degrees
Muscovite φ′r = 17 to 24 degrees
Hydrous mica φ′r = 16 to 26 degrees
Quartz-dune sand φ′r = 30 degrees
Crushed sand φ′r = 35 degrees
Appendix B Page B2

Table B1.2: Shear strength on relict discontinuities in weathered rocks in Hong Kong
(Irfan and Woods 1998; Irfan 1997)
Type Nature Strength Method of Remarks
Parameters Determination
Discontinuity Completely decomposed φ′= 37 °
Direct shear
Surface planar joint
Completely decomposed φ′= 32° Direct shear
planar joint with MnO2
and FeO2 coating
Highly decomposed φ′= 45 to 57° Not known Peak (?)
rough joint in granitic
saprolite
Highly decomposed clean φ′= 38° Direct shear Portable shear box
joint in granitic saprolite c′ = 30 kPa
Highly decomposed joints φ′= 40 to 42° Direct shear Basic frictional
in granite strengths from lab tests.
Granitic saprolite Not known Compiled laboratory
- clay coated joint φ′= 25° ± 2 ° results by Lumb (1971)
- clean joint φ′= 35° ± 2 °
Completely decomposed φ′ = 29° Direct shear Failed slope, Pun Shan
joint in volcanic saprolite c′ = 22 kPa Tsuen.
Stained or clean joint in φ′= 23 to 42° Multi-stage
volcanic saprolite direct shear
Highly weathered φ′ = 38° CU triaxial Failed slope at Shum
volcanic rock c′ = 5 kPa Wan Road.
Completely decomposed φ′ = 37° CD triaxial,
tuff, intact c′ = 8 kPa multi-stage and
direct shear
Completely decomposed φ′ = 29° Multi-stage
rough foliation plane in (intact material direct shear
meta-sandstone with φ′= 34°, c′ =3.5kPa)
phyllite
Rhyolitic saprolite Not known Compiled laboratory
- clay coated joint φ′= 22.5 ° ± 2° results by Lumb (1971)
- clean joint φ′= 30° ± 2 °
Discontinuity Completely decomposed φ′ = 34° Direct shear Failed slope at Sui Sai
with Infilling discontinuity with kaolin Wan (kaolin infill may
infill not be fully saturated;
alignment problems in
shear box)
Completely decomposed φ′ = 36° CU Triaxial Lower failure surface,
discontinuity in granitic Tsuen Wan
saprolite
Completely decomposed φ′ = 31.5 ° CU Triaxial Upper failure surface,
discontinuity in granitic Tsuen Wan
saprolite
Joint filled with stiff φ′ = 31° Direct shear Portable shear box
brown clay in granite c′ = 30 kPa
Kaolin-filled joint in φ′ = 26° Direct shear Portable shear box
granite
Joint with 20 mm thick φ′= 27.5 °, c′ = 31kPa Back analysis Failed slope at Junk
kaolin-infill along rough (kaolin infill: (direct shear) Bay Road.
discontinuity in highly to φ′=24.5 °, c′ =243kPa) (full saturation of
completely weathered kaolin may not been
granite achieved)
Clay filled joint in φ′ = 22° Multi-stage
volcanic saprolite direct shear
Appendix B Page B3

Table B1.2 (continued): Shear strength on relict discontinuities in weathered rocks in


Hong Kong (Irfan and Woods 1998; Irfan 1997)
Type Nature Strength Method of Remarks
Parameters Determination
Discontinuity Manganese coated joint in φ′ = 14 to 26° Direct shear For slickensided and
with Infilling meta-volcanic saprolite (20 ° average) non-slickensided.
(Intact completely
decomposed meta- (φ′ = 32 °, c′ = 6 kPa) CU triaxial
volcanic)
Slickensided joint in with φ′r = 16° Direct shear Failed slope in Brazil
smectite infill in migmatic
saprolite
Joint with black seam in a φ′ = 14.5 °, c′ = 10 kPa CU Triaxial
gneissic saprolite
(Slickensided) φ′r = 10.5 °
(Intact, φ′ = 18.5 °)
Shear Zone 0.7 m thick fault zone φ′ = 32 to 39° CD triaxial,
(mixture fractured rock c′ = 0 to 14 kPa multi-stage and
with kaolin infilled joints direct shear
in moderate to highly
decomposed volcanic)
Slickensided, pre-existing φ′ = 18° CD direct shear Failed slope sat
shear surface between Borrow Area A, Tuen
volcanic saprolite and Mun (possible
colluvium difficulties with
aligning shear surface)
Pre-existing shear surface φ′ = 34°, c′ = 3.5 kPa Direct shear Failed slope at Island
with up to 30 mm thick (intact completely Road Government
iron rich clay seam decomposed tuff, School
between creeping and φ′ = 30 to 33°
insitu volcanic saprolite c′ = 12 to 22 kPa)
(Creeping layer,
φ′ = 30°, c′ = 6 kPa)
0.6 m thick, completely φ′ = 29° Direct shear Failed slope at Fei
decomposed, altered tuff (φ′ = 22 ° lower bound) Tsui Road
with kaolin clay
veins/infilling between (intact kaolinised
less weathered rock material, φ′ = 22°)
(altered bedding shear
surface)
Clay infilled joints in Clay seam, φ′ = 26°, CU triaxial Failed slope at Shum
highly weathered volcanic c′ = 8 kPa Wan Road
rock Slickensided, φ′ = 21 ° Direct shear
Jointed Completely weathered φ′ = 30° Probabilistic Failed slope at Area
Saprolite meta-volcanic saprolite c′ = 5 kPa analysis 19, Tuen Mun
Highly weathered φ′ = 11 to 12° Back analysis Failed slope at Chai
(Mass) volcanic saprolite c′ = 32 to 36 kPa Wan (absence of
accurate groundwater
data)
Completely weathered φ′ = 34°, c′ = 2 kPa Back analysis Failed slope at Sui Sai
volcanic tuff saprolite Wan
with multiple shear (failed mass,
surfaces φ′ =28.5 °)
Appendix B Page B4

An explanation of some of the terminology used in Table B1.3 and Table B1.4 is as
follows:
• Slope Type; ‘Cut’ for cut slope, ‘Fill’ for fill slope and ‘Nat’ for natural slope.
Retained slopes are identified with an R following the type of slope retained; e.g.
Fill-R refers to a retained fill slope.
• The definitions for slope failure geometry are given in Section 1.4 of Appendix A.
1 = Type 1; 2 = Type 2; 3 = Type 3 and 4 = Type 4.
• Initial slide classification. Several additional terms used that are not defined in
Section 1.2 of Appendix A are:
− Interface - the surface of rupture is coincident with an interface between
different material types, such as colluvium and saprolite or fill and natural
ground.
− Defect control – defects have a strong influence on the location of the surface of
rupture.
• Travel classification. Classification of the post-failure movement of the landslide.
Refer to Section 1.2 of Appendix A for definitions of the terms used.
• Obstructed. Consideration has been given to whether or not the travel of the slide
mass has been obstructed. For example, in the case of Sau Mau Ping (1976) the
travel distance of the slide was clearly obstructed by the building at the toe of the fill
slope.
• Landslide Geometry. The terms used are defined in Section 1.3 of Appendix A.
• The remainder of the terminology used is self explanatory; however, some terms or
acronyms have been used that require further explanation:
− CW – completely weathered rock
− PW – partially weathered rock
− SM, ML, MI, CL, CI, CH – soil classification symbols of the Australian soil
classification system (Australian Standard AS 1726-1993 “Geotechnical Site
Investigations”).
− SMDD – Standard Maximum Dry Density as determined by the Standard
compaction test (Australian Standard AS1289.5.1.1.).
− SPT – Standard penetration test
− DAN – numerical computer program DAN
− phiB – bulk friction angle used for frictional model analysis in DAN.
Appendix B Page B5

For the flow slide failures in loose fill slopes in Hong Kong, assessment of the
potential for the fill material to be contractive on shearing has been based firstly on
measured density results (if available) and then on the type of placement and age of the
fill.

1.2 “RAPID ” LANDSLIDES FROM FAILURES IN GRANULAR FILLS

Virtually all of the landslides in mine waste spoil piles and stockpiled coking coal
within the database are classified as flow slides of “rapid” post failure velocity. The
granular fills within which the failures occurred are classified as predominantly sandy
gravels and gravely sands with trace to some silt fines (up to 12% silt fines). Data for
some 67 “rapid” landslides in loose, granular stockpiles of coal mine waste and coking
coal have been collected and analysed. They come from three geographic regions, all
associated with the coal industry, summarised as follows:
• Landslides in loose, stockpiled coal mine waste from South Wales (18 cases);
• Landslides in loose, stockpiled coal mine waste from British Columbia, Canada (40
cases); and
• Landslides in loose, stockpiled coking coal from Hay Point, Australia (9 cases).

A summary of the collected information of each case study is given in Table B1.5.
The regional designation of case study is identified by the prefix to the case study
number; SW for South Wales, BC for British Columbia, and HP for Hay Point.
For most case studies only limited information has been collected. For the British
Columbian cases the information for most case studies has been obtained from Golder
Assoc (1995) supplemented with more detailed information on three cases from
Dawson (1994) and Dawson et al (1998). For South Wales most of the case study
information has been sourced from Siddle et al (1996) supplemented with additional
information on the Aberfan (1944, 1963 and 1966 events) and Cilfynydd (1939) flow
slides from the Aberfan tribunal report (H.M.S.O. 1967) and associated technical
reports (Bishop et al 1969; Bleasdale 1969; Woodlands 1969). For the Hay Point case
studies the available information on the flow slides in coking coal stockpiles at the
loading facility in North Queensland were sourced from Eckersley (1985, 1986).
In most cases the landslides have been classified as flow slides, where the failure
mechanism and rapid acceleration of the failed mass is associated with contraction and
Appendix B Page B6

liquefaction of the loose, stockpiled material (in a saturated or near saturated condition)
on shearing. The travel of the slide mass, in most cases, has been classified as a flow
slide. In a number of the observed failures non-liquefiable material has been carried
forward on a basal liquefied layer (non-liquefiable because of the degree of saturation or
high permeability, as in the case of the coarser sandstone waste in the British Columbia
spoil piles). In some cases it is possible that the flow slide develops into a turbulent
debris flow.
Comments on specific issues relating to each sub-group of flow slides in coarse,
granular materials are as follows:
• For South Wales, the initial slide classification in most cases, as described by Siddle
et al (1996), is rotational. It is considered likely; however, that each slide has a
basal translational component along the fill / foundation interface and an arcuate
back-scarp through the spoil pile material. Apart from several slides, the volume of
the initial slide that developed into a flow slide is relatively low (less than about
50%) with a significant portion of the initial slide volume retained in the failure
bowl. Estimates of both the initial slide volume and the volume involved in the
flow slide are given in Table B1.5. For the failure of Tip 7 at Aberfan in October
1966, Bishop et al (1969) considered it likely that the failure occurred as several
retrogressive components in rapid succession, that initiated from a localised failure
in the toe region (partly along a pre-existing shear surface) triggered by high
foundation pore water pressures in this region. It is possible that a similar slide
mechanism occurred at a number of the other failures given the number of
landslides in tips in South Wales located over pre-existing earthflows and spring or
stream lines.
• For the coalfields in British Columbia, Golder Assoc (1995) indicates that the initial
landslide is dominantly translational. They describe the typical slide as having a bi-
linear profile with a basal component close to the fill / foundation contact and a
steep back-scarp (45 to 55 degrees). The parallel layering of the liquefaction
susceptible finer zones of sandy gravel within the very coarse cobbly and bouldery
sandstone waste material would also be expected to influence the shape of the
surface of rupture.
• At Hay Point, the initial landslide classification is considered to have a significant
translational component along the fill / platform interface. Observation of an actual
event (Eckersley 1986) and the large scale laboratory tests performed by Eckersley
Appendix B Page B7

(1986) indicate it is likely that these flow slides develop as a series of retrogressive
failures initiated in the toe region of the stockpile. Eckersley indicates that the bulk
of the material in the slide mass is carried forward on a liquefied basal layer of the
saturated coking coal.

The slide volumes for the most of land slides from South Wales and Hay Point were
estimated based on the geometric details provided for each slide. For the South Wales
slides the volume of the initial failure that transformed into a flow slide, where it was
not reported in the published literature, was based on estimation from the information
provided. For a number of cases the estimate is very rough.

1.3 STRONGLY RETROGRESSIVE LANDSLIDES

The groups of case studies considered as strongly retrogressive include flow slides from
failures tailings dams, sensitive clays, submarine slopes and sub-aqueous hydraulic fills.
The details of the case studies used in the analysis from these slide groups are
summarised in the following tables:
• Table B1.6, flow slides from tailings impoundments initiated by static liquefaction;
• Table B1.7 flow slides from tailings impoundments initiated by dynamic
liquefaction; i.e. earthquake triggered flow slides;
• Table B1.8 flow slides in sensitive clays; and
• Table B1.9 landslides in submarine slopes.

The case studies of sub-aqueous fills are given in Table B1.10 with the hydraulic fill
embankment dam case studies.

1.4 “RAPID ” LANDSLIDES FROM FAILURES IN HYDRAULIC FILL EMBANKMENT


DAMS
The five case studies of flow slides include four failures in hydraulic fill embankment
dams (Fort Peck, Calaveras, Lower San Fernando and Sheffield dams) and one failure
(Wachusett dam) in the upstream shoulder of a loose dumped sand fill embankment that
failed during first filling. Olson et al (2000) considered the failure of Wachusett Dam to
Appendix B Page B8

have involved static liquefaction and flow of the loose sand fill. The details for each
case study are summarised in Table B1.10.
Of the four hydraulic fill embankment dams, the initial failure was associated with
liquefaction of either the hydraulically placed fill or the foundation. Two of these
failures (Fort Peck and Calaveras dams) occurred during construction and the other two
(Lower San Fernando and Sheffield dams) shortly following an earthquake. An
important characteristic of these failures is the stratified nature of the hydraulically
placed materials. A noted observation of the post failure travel from the published
literature it that it was evident that the outer denser materials and non-saturated zones
remained virtually intact, being carried forward on a liquefied zone of the looser and
saturated materials.
Appendix B Page B9

Table B1.3: Landslide case studies from Hong Kong


Landslide Geometry Width (m)
Slope Travel
Date of Slope Initial Slide Travel Obstr- Confined Slope Slide Slide Ratio Volume Material Description
No. Site Failure Length, Failure Alpha 1 Alpha 2 Alpha 3 Alpha 4 Distance 3 Initial Geology Hydro-geological Features Morphological Features
Failure Type Classification Classification ucted (Yes/No) Height, Height, Length, L Depth, D H/L (m ) Runout (Slide Source)
Geometry Ltoe (m) (deg.) (deg.) (deg.) (deg.) Angle (o) Slide
Hs (m) H (m) (m) (m)
Underlying weathered rock less permeable than colluvium. Steep cut (51 degrees) in colluvium and
Cheung Shan Slide of debris,
HK1 16/6/93 Cut 1 debris flow Yes No > 25 17.7 24 14.6 1.4 51 51 0 na 36.4 0.74 35 9 8 Granodiorite Colluvium - loose, permeable. Infiltration resulted in saturation of colluvium and transient PW granodiorite. Natural slope above cut
Estate translational on interface
perched water table at colluvium / PW granodiorite interface. of 36.5 degrees.

Evidence of surface seepage on slope 1986 & 1987.


Residual, saprolitic soils and PW rock. Relict Seepage observed flowing out of failure scar (high or Chunam covered cut slope (41 degrees) in
Compound - defect
HK2 Allway Gardens 9/27/93 Cut 1 debris flow No No 54 35 55.2 40.2 2.3 41 41 0 na 32.4 0.63 225 12 12 to 15 Volcanics joints coated with 10 mm layer of soft to firm perched groundwater). Suspect rapid variation in residual soil, saprolite and PW volcanic
control
clay observed in landslip scar. groundwater following rainfall and rain filled tension rock. Steep natural slope above cut.
cracks. Catch drain upslope of failure (at top of cut).

Kau Wa Keng Residual soil (loose to medium dense, clayey Springline upslope of failure. Infiltration of seepage and
Compound - significant (30 slope Very steep cut slope (up to 70 degrees) in
San Tseun - debris slide / debris silty gravelly sand) over saprolite over PW rainfall. Development of transient perched water table
HK3 4/6/97 Cut 1 translational component, Yes No 80 9.8 12.6 8.4 2 to 3 70 45 0 above 37.9 0.78 80 8.5 9 Granite residual soils and saprolite. Chunam
Slide flow granite. Slide exposed closely spaced above weathered rock. Several erosion pipe holes
defect control cut) covered cut slope with weepholes.
A discontiuities parallel to to slope. observed in scarp.

Infiltration of surface water during heavy rain saturating


Keng Hau Road - Compound - non circular, Residual soil - clayey silty sand overlain by Steep cut slope (53 degrees) in residual
HK4 2/6/97 Cut 1 debris flow No No 31.5 31.5 64.5 39 5 53 40 6 na 26.0 0.49 210 16 14 to 7 Granite slope, possible development perched water above
Slide B retrogressive shallow fill (granular and porous). soil and saprolite.
saprolitic soils.

Saprolite - moist to wet silty and sandy clay,


high plasticity and soft to stiff (SPT < 3). Possible development of high perched groundwater
Hong Tseun Cut slope (36 to 42.5 degrees) in saprolitic
HK5 3/7/97 Cut 1 Compound - non circular debris flow No No 15.3 15.3 33.3 23.3 2.5 36 42.5 0 na 24.7 0.46 250 25 18 to 8 Volcanics - Tuff Underlain by stiff to very stiff silty sandy clays, pressures above stiff clays due to infiltration from rainfall.
Road soils.
significant gravel to boulders. Jointing, not Seepage observed coming out of scarp.
adverse to failure.

Chung Shan Saprolitic soil - medium dense silty fine to Groundwater table well below failure. Cut off drains at top Steep cut slope (52 degrees) in saprolitic
debris slide ? / debris 50
HK6 Terrace - Slide 4/6/97 Cut 1 Compound - non circular Yes No 18.5 12.4 14.3 5.6 3 52 0 na 40.9 0.87 85 16 14 Granite coarse sand. No indication of unusually and toe of cut slope. Possible build up of perched water soils. Unprotected. Steep, sparsely
flow ? (0 to 90)
B weak materials or adverse discontinuities. ollowing infiltration. vegetated natural slope above cut.

Steep cut (60 to 65 degrees) in saprolite


Saprolite (weathered tuff in matrix of clayey Inflitration of rainfall into slope. Development of perched
and PW rock. Upper part of cut chunam
Translational - defect silty sand / sandy silt) of 4 to 11 m depth water table on altered tuff (evidenced by seepage). Surface
HK7 Fei Tsui Road 13/8/95 Cut 1 debris flow, dry No No 27 34 72 34 15 65 65 -2 na 25.3 0.47 14000 40 to 70 95 Volcanic - Tuff covered. 12 degree natural slope above
control overlying PW rock. Persistent relict jointing drains at crest of cut and on slope suspected of being
cut with reservoir some 30 m back from
in saprolite and altered tuff layer (ML/MI). blocked.
cut.

Saprolite (slightly clayey sand) over PW rock. Groundwater below base of failure. Blocked surface Steep cut slope (52 to 75 degrees)
Ten Thousand Persistent sheeting joint in conjunction with drains at crest. Rapid infiltration through open joints and covered with deteriorated chunam
debris slide / debris 62
HK8 Buddha's 2/7/97 Cut 1 Compound - defect control Yes No 18 27.6 39.6 20.1 5.5 52 0 na 34.9 0.70 1500 25 42 Granite other joint sets were adverse to stability of the erosion pipes in saprolite resulted in development of (severely cracked and detached in places).
flow (56 to 75)
Monastery cut. Overlain by shallow layer of permeable perched water table above PW rock. Seepage observed Densely vegetated natural slope (36
colluvium. in scarp post slide. degrees) above cut.

Cut slope (48 degrees) of 27 m height


Saprolite - gravelly silty and clayey sand. Subsurface transient groundwater controlled by a system in 4 benches. Surface drains installed.
Ching Cheung
HK9 3/8/97 Cut 1 Compound - defect control debris flow No No 27.4 31.9 81 51 2 to 4 48 38 0 na 21.5 0.39 2000 35 32 to 43 Granite Indications of possible adverse jointing. of natural pipes and basalt dyke aquitards. Raking drains in Slope mostly vegetated, some
Road
Overlain by shallow layer of colluvium. lower two batters. Seepage in scarp observed post failure. shotcreting. Natural slope above cut
approx 31 degrees.

Development of transient perched water table in soil Steep cut slope (52 degrees) in
Fung Wong Saprolite - gravelly silty sand. Adverse, partly
Translational - wedge mass above relict discontinuities. No control of surface decomposed granite. Shotcrete and
HK10 Service 9/6/98 Cut 1 debris flow, dry Yes No 26.5 29 29.3 15.7 2 to 2.5 52 51 0 na 44.7 0.99 120 8.5 5 to 9 Granite kaolin infilled relict joints (10 to 20 mm
type, defect control water from above (blocked drain at crest), therefore chunam covered. Natural slope of
Reservoir thickness) in soil mass.
exaserbating infiltration. about 36 degrees above cut.

Steep cut slope (50 to 60 degrees), old


Saprolite - moist, medium dense slightly Inadequate surface water drains. Erosion pipes in soil
Yue Sun Garden and over-steep, with no slope protection.
HK11 9/6/98 Cut 1 Compound - non circular Debris flow No No 9.2 10.2 14.2 8.5 1.2 56 56 9.5 na 35.7 0.72 200 13.5 13.5 Volcanics - Tuff clayey sandy silt with some gravel (scarp mass provided preferential sub-surface flow paths. No
- Slide A Natural slope above cut 16 to 20
description). seepage observed 2 days after slide.
degrees (densely vegetated).

Residual soil - clayey silty sand overlain by Infiltration of surface water during heavy rain saturating
Keng Hau Road Steep cut slope (64 degrees) in residual
HK12 4/6/97 Cut 2 Compound debris flow No No 11 29 83 66.5 5 64 19.5 6 na 19.3 0.35 150 10 to 12 10 to 12 Granite shallow fill (porous granular fill) of 0.7 m slope, possible development of perched water above
- Slide A soil.
thickness. saprolitic soils.

Concentration of surface and sub-surface water flow Benched cut slope, average 55 degrees,
Ha Wo Che, Compound - non circular Saprolitic soil - closely spaced discontinuities
HK13 3/7/97 Cut 2 debris slide Yes No 9.3 8.8 12.9 6.4 2.5 55 0 - na 34.3 0.68 60 10 14 Granite directed toward landslip site. Possible build up of covered with chunam. Natural slope of
Shatin (some defect control) oriented sub-parallel to slope.
pore pressures in slope. 25 to 40 degrees above cut.

Groundwater table below base of failure. Likely water ingress


St Joseph's, Compound - non circular Rubble facing on steep cut slope (65
HK14 3/8/97 Cut 2 debris flow, dry Yes No 8.6 7.8 8.2 3.6 1 65 0 - na 43.6 0.95 25 3 to 5.5 8.2 Granite Weathered granite (saprolite ?) due to overflow from blocked drain at crest. Possible delay
Ngau Tau Kok (some defect control) degrees) in weathered granite.
between rainfall and failure.

Likely water infiltration at crest of cut slope and development Stone facing on steep cut slope (66 degrees)
Bayview
HK15 2/7/97 Cut 2 Rotational debris flow No No 14.5 13 12.1 6 1 to 2 66 0 - na 47.1 1.07 70 11 11 Volcanics - Tuff Saprolitic soil - silty fine to medium sand. of perched water behind rubble wall (blocked weep holes). in volcanic saprolitic soils. Densely
Gardens Seepage observed from failure scar. vegetated 5 degree slope above cut.

Cut slope (50 to 60 degrees), old


Inadequate surface water drains. Erosion pipes in soil
Yue Sun Garden Compound - significant debris slide / debris 52 Saprolite - firm to stiff sandy clayey and over-steep, no slope protection.
HK16 9/6/98 Cut 2 No No 14.1 15.5 27.7 13.2 3.5 2 - na 29.2 0.56 1100 36 28 to 17 Volcanics - Tuff mass provided preferential sub-surface flow paths. No
- Slide B translational component flow (50 to 60) silt (scarp description). Relict joints. Natural slope above cut 16 to 20
seepage observed 2 days after slide.
degrees (densely vegetated).

Groundwater below base of slide. Surface drainage on cut


Slide of debris, Colluvium (loose fine to coarse grained sand
Lai Cho Road, slope and catch drain at crest. No seepage observed. Shallow cut (34 degrees) in colluvium and
HK17 4/6/97 Cut 3 compound - non circular debris flow No No 54 28.2 44.4 34.4 1 34 34 - na 32.4 0.64 33 6 6 to 7 Granite with some boulders). Underlain by saprolitic
Kwai Chung Direct infiltration of rainfall on slope and potential saprolitic soils. Flat slope above cut.
(on interface ?) soils (exposed in landslide scar).
development of perched water table.

(23 Blocked drain above slide, overflowed onto cut slope.


Castle Peak Granite. Intruded by a Saprolitic soils - gravelly silty sand with
natural Basalt dykes controlled groundwater regime (dammed After initial failure (July 1994) cut
HK18 Road (Milestone 7/8/94 Cut 4 Rotational debris flow, dry No No 9.1 24.5 45.8 28.9 5.5 46 0 30 28.1 0.53 300 30 22 number of sub-vertical cobbles and boulders. Near vertical basalt
slope groundwater flow and raised groundwater level) within slope. slope cut back from 60 to 46 degrees.
14.5) basalt dykes (CW). dykes (CW).
above cut) Water seepage within landslide scar uphill of dykes.

17 (24 Saprolitic soils (clayey silty sand with gravel) Infiltration from rainfall and surface water. Development of
debris slide / debris 14.5 m high cut (57 degrees) at toe of
HK19 Po Shan Road 18/6/72 Cut 4 Compound - defect control No No 14.5 120 270 175 6 to 16 57 0 reducing na 24.0 0.44 40000 25 to 40 60 to 70 Volcanics to 10 to 30 m. Overlain by colluvium (silty clay perched water tables (possible rise in groundwater ?).
flow steep (30 to 40 degrees) natural hill slope.
to 13) to clayey silty sand) of approx 3 m depth. Seepage from toe of cut and upslope of cut (spring).

Transient perched groundwater developed in weathered Low height (4.5 m), steep cut (60
Saprolite (granitic). Persistent 20 mm thick
60 33.5 granite on discontinuities following heavy rainfall. degrees) on natural slope (31 to 33.5 deg).
Kau Wa Keng Translational - defect joint formed surface of rupture. Overlain by
HK20 4/6/97 Cut 4 debris flow No No 4.4 29 48.8 29.8 1 to 5 (31 above (some 0 na 30.7 0.59 280 14 8 to 18 Granite Infiltration through open discontinuities. Seepage Run-out predominantly on natural slope
Upper Village control shallow layer of permeable colluvium (1 m
cut) 9 and 65) observed from main scarp above discontinuity, which (below cut) with some flatter and steeper
depth).
reduced with time after cessation of rainfall. slopes associated with benching.

Thin masonry block retaining wall


Saprolite - sandy silt to silty fine sand,
3.5 Leakage from defective stormwater and sewer services (at 75 deg slope) supporting cut in
HK21 Kwun Lung Lau 23/7/94 Cut - R 2 Compound (retained slope) debris flow No No 9.2 12.3 23.6 15.3 3 to 6 75 - na 27.5 0.52 1000 28 33 to 38 Volcanics - Tuff completely to highly weathered tuff, very wet.
(0 to 90) saturated retained soil. Small contribution from rainfall. saprolitic soils. Fill placed above
Overlain by loose fill, placed 1964 to 1969.
retaining wall, chunam covered.
Appendix B Page B10

Table B1.3: Landslide case studies from Hong Kong (Sheet 2 of 4)

Slope Post-Failure Characteristics Velocity


No. Site Rainfall Pre-Failure Observations Trigger Mechanism Initial Slide Description Post-Failure Observations Comments References
Type (Obstructions, deposition, channelisation) (m/s)

Wet period for 15 to 20 days prior to slide. Rainfall (saturation of colluvium Instability due to rise in groundwater Slide of debris from top of cut slope. Bus and bus shelter obstructed slide at toe. No
Cheung Shan Very quick flow. Rapidity due to angle of cut slope (51
HK1 Cut Heavy rain in 1.5 hrs leading up to slide, na and development of perched leading to translational debris slide on Translational - on interface between channelisation. No deposition on 51 deg slope, Failure entirely within colluvium. One person killed. Chan et al (1996)
Estate deg).
peak 79 mm/hr (1 in 10 year event). water table). colluvium / weathered rock interface. colluvium and PW granodiorite. deposition on flat slope at toe.

Sept 93 generally dry except for 2 periods New chunam placed in 1984 at approx. location
Instability due to groundwater and water No channelisation. Minor obstruction (edge of
of rain - 125 mm (13 to 18/9/93) and 390 of the slide, therefore possibly some minor
Rainfall (likely rise in filled tension cracks. Consider also time Defect controlled compound slide in building on cut slope). No deposition on steep cut Cut slope covered with chunam in 1984.
HK2 Allway Gardens Cut mm (23 to 27/9/93). 1 in 5 yr event for problems at this time. No prior signs of failure Loud noise followed by release of debris (soil and rock). na Chan et al (1996)
groundwater level in slope). effects of softening of relict joints and partially weathered rock. slope (41 deg), deposition on flat slope at toe (max Constructed between 1976 to 1978.
25 & 26/9/93. Little rainfall on 27/9/93 leading up to slide. Pre-existing tension cracks
strain weakening. 3m)
prior to slide. on upper part of failure scar, up to 2 m depth.

Kau Wa Keng Defect controlled compound slide in


Minor landslide in base of cut slope on 8/5/97 Rainfall (infiltration into soil Instability due to reduced strength of Part of slide mass retained in failure bowl in relatively
San Tseun - 24 hr - 263 mm, 12 hr - 252.5 mm, max residual and saprolitic soils. Significant Obstructed by house. No channelisation. Deposition
HK3 Cut during heavy rain (one month prior to this and development of perched saturated soil and perched groundwater intact condition. Debris that exited source area came to na Constructed before 1954. No washout effects. Halcrow (1999a)
Slide intensity 128.5 mm/hr (1 in 40 yr event). translational component along influenced by house.
failure). water table). weathered rock. rest against rear wall of house. No influence of washout.
A discontinuities parallel to face.
Compound (non-circular) slide through
Heavy rainfall for 4 hours prior to slide Large failure preceeded by a smaller failure. Two distinct failures. Initial small failure that was highly
Keng Hau Road - Rainfall (infiltration through Instability due to loss of strength on residual soil mass, possibly translational Deposition on full length of slope - 2 to 4 m deep at Loss of lateral support from Slide A possibly a
HK4 Cut (max. 1 hour = 97.5 mm/hr). 1 in 34 Possibly only short time period between mobile followed by a second larger failure of lesser na Halcrow (1999e)
Slide B platform above cut slope). saturation and possible perched water. at toe on saprolite interface. toe of cut slope, much shallower at distal end. significant factor.
year event for 2 hr rainfall. failures. mobility.
Retrogressive.

Slide (compound, non-circular) through No obstruction or channelisation. Most of


Rainfall (infiltration and
Hong Tseun Heavy and prolonged rainfall. 31 day - Instability associated with saturation soil mass in saprolitic soil. Within zone of deposition on flat slope at toe of cut (2 to 0.5 m), Construction completed 1978. Travel down
HK5 Cut na development of perched na na Halcrow (1999d)
Road 1193 mm (1 in 40 yr event). of soils and perched water. soft to stiff clay, base close to boundary remainder above first berm. No deposition on road due to washout. Fully vegetated by 1997.
groundwater table).
with very stiff clay. 42.5 deg slope.

Chung Shan Rainfall (infiltration and No channelisation. Obstruction from wall in front of Slope upgraded in 1990. History of minor
Instability associated with saturation Failure through saprolitic soil mass,
HK6 Terrace - Slide Cut Heavy rainfall - 4 hr event 1 in 50 years. na development of perched na slide. Deposition on lower portion of scarp and flat na failures. Slope vegetated with small shrubs and Halcrow (1999d)
of soils and possible perched water. compound (non-circular). Shallow.
B groundwater table). slope at toe (2 to 3 m depth). grass.

Instability due to rising perched water Hungr GR (1998)


Translational defect controlled failure in DAN model - good representation of deposition using
Heavy prolonged rainfall. 31 day rainfall Rainfall (infiltration leading to table from prolonged and heavy rainfall. Main failure preceeded by small failure in upper cut slope. No channelisation or obstruction. Deposition in Irfan (1997)
saprolitic soil mass. Altered tuff layer 10 to 13 phiB = 24 deg on rupture surface and 36 deg on distal
HK7 Fei Tsui Road Cut - 1303 mm (1 in 95 year event), peak na rise in perched water table Likely slide brittleness associated with Rapid movement of main slide (slid down in a couple of bowl of failure (4 m deep on 15 deg slope) and on
(DAN) path. GEO consider low travel angle possibly due to
Irfan & Woods (1998)
formed base of slide, persistent joints
94.5 mm/hr in hours prior to failure. above altered tuff layer). strain weakening along controlling seconds). Large degree of break up of slide mass. flat slope at toe (10 m max at toe of cut). GEO (1996a)
formed backscarp. angle of trajectory of failed mass.
defects and lateral margins. Knill (1996a)

Instability due to high perched Defect controlled compound slide in


Ten Thousand Heavy rainfall on morning prior to slide. Rainfall (infiltration and water table from heavy rainfall. saprolite. Persistent sheet joint in central Failure occurred suddenly during torrential rain. Fast No channelisation. Obstructed by building at toe of
Defect controlled failure in saprolitic soils and
HK8 Buddha's Cut 1 in 25 year for 1 hour prior to slide (118.5 Open joints evidence of past movement. development of perched Brittleness associated with strain portion and discontinuous clay joint in moving (quickly buried annex of building). Significant cut slope. Deposition - 4 m deep in failure bowl on 20 na Halcrow (1998b)
weathered rock.
Monastery mm/hr). groundwater table). weakening along controlling defects lower part. Steeply curved degree of break up at toe, lesser back in scarp. deg slope), up to 9 m deep on flat slope.
and internal slide brittleness. backscarp.

History of failures in this slope. Failure on


Failure on 7/7/97 occurred 2 days after Instability due to development of perched
3/8/97 (this failure) preceeded by two failure in No channelisation or obstruction. Deposition on flat Defect controlled landslide in granitic saprolite
heavy and prolonged rainfall (31 day 1 Rainfall (infiltration and water table and gradual weakening of slope
Ching Cheung July 1997. Extensive area of cracking and Defect controlled compound slide in slope at toe (1 to 2.5 m depth). Mobility of slide soils. Cut formed in 1967. History of numerous
HK9 Cut in 500 yr event). Failure on 3/8/97 development of perched from progressive movements. Suspect Run out distance suggests the slide was quite mobile. na Halcrow (1998a)
Road displacement at top of slope (large perimeter saprolitic soil. possibly enhanced from ponded water on enclosed failures. Possible weakening of slide mass in
preceeded by 278 mm in 3 days (1in 2 groundwater table). brittleness associated with internal slide
crack 2.5 m vert settlement, 1 m horiz) section of Ching Cheung Road. period up to slide on 3/8/97.
year event), 31 day event 1 in 49 years. mechanism.
observed prior to failure.

Defect controlled translational wedge type


Fung Wong Instability associated with build up of Sinlge phase rapid event. Some debris flowed into Obstructed by fence around edge of reservoir.
Heavy rainfall in 4 hours prior to failure, failure in saprolite along persistent Cut slope formed before 1963. Debris
HK10 Service Cut na Rainfall perched water. Partly controlled by reservoir. Loose silty sand material with gravel, dry at Partial channelisation. Deposition mainly on flat na Fugro (1999a)
2 hour rainfall 1 in 8 year event. discontinuity. Shallow. Part failure observed to be dry.
Reservoir persistent relict discontinuity. time of inspection. surface at toe, some spilled into the reservoir.
through soil mass.
Instability associated with infiltration
Rainfall (infiltration and
Heavy rainfall. 2 to 12 hour events, leading to wetting up and seepage Compound (non-circular) failure through
Yue Sun Garden development of seepage No channelisation or obstruction. Deposition on
HK11 Cut 1 in 5 to 6 year return period. Peak History of small failures in this slope. pressures. Progressive weakening saprolitic soil mass. Shallow spoon na na Cut slope formed 1961. Fugro (1999d)
- Slide A pressures along preferential flat slopes below toe of cut, 3 m max at toe of cut.
hourly rate = 60.5 mm/hr. with time of old cut slope also a shaped surficial slip.
flow paths).
contributing factor.
Heavy rainfall for 9 hours prior to slide,
Keng Hau Road Rainfall (infiltration through Instability due to loss of strength on Compound slide through residual soil Highly mobile slide, possibly due to stormwater discharge Deposition on full length of slope, 0.7 m near rupture Poor condition of slope surfacing. High
HK12 Cut > 305 mm (max. 1 hour = 81.5 mm/hr). na na Halcrow (1999e)
- Slide A platform above cut slope). saturation and possible perched water. mass. on to slope from drainage channel near toe of scarp. and 0.1 to 0.3 m at distal end. mobility of slide mass.
1 in 22 year event for 4 hr rainfall.
2/7/97 - heay rain preceeded slide (1 in
3 Compound (non-circular) failure through No channelisation. Slide obstructed by building at
Ha Wo Che, 40 yr for 1 to 4 hr duration) Preceeded by small failure (5m ) on previous Instability associated with wetting up Constructed in early 1980s. Evidence of
HK13 Cut Rainfall saprolitic soil mass, some control by Evidence of displacement along discontinuities. toe. Slide only moved some 2 m at head of scarp, na Halcrow (1999e)
Shatin 3/7/97 - heavy prolonged rainfall (1 in day and possible perched water. displacement along discontinuities.
discontinuities. most of deposition remained within failure bowl.
120 yr event for 1 to 31 day)
Significant obstruction. Slope has a history of
Landslide followed 2 periods of wet Rainfall (infiltration at crest Instability associated with saturation and Compound (non-circular) shallow slide in
St Joseph's, Obstruction by building at toe of slide. No past failures. Heavier rainfall in past,
HK14 Cut weather in July. Prior to slide, rainfall 1 in 1 na and open joints in facing build up of water pressures behind rubble weathered granitic saprolite. Slide scarp na na Halcrow (1999c)
Ngau Tau Kok channelisation. Deposition at toe, max 2 m depth. deterioration of facing likely to have had an
to 3 for all return periods. blocks). facing. oriented to one of the major joint sets.
influence.

Heavy rainfall for 8 hours prior to slide. 31 Instability associated with saturation and Cut - pre 1977. Rainstorm not particularly severe
Bayview Loud noise heard at time of slide. Some washing out of No obstruction. No channelisation. Deposition at toe
HK15 Cut day rainfall 879.5 mm. Peak hourly = 72.5 na Rainfall build up of water pressures behind rubble Rotational - shallow. na therefore deterioration of slope and blockage of weep Halcrow (1999c)
Gardens debris by post failure surface water flow. of slope, max 2 m depth.
mm/hr (1 in 6 yr event). facing. holes probably had an effect on stability.

Instability associated with infiltration


Rainfall (infiltration and
Heavy rainfall - 12 to 24 hour events, leading to wetting up and seepage No channelisation or obstruction. Deposition
Yue Sun Garden development of seepage Compound - significant translational Debris in slide remained relatively intact, relatively quick Cut slope formed 1961. Slide debris
HK16 Cut 1 in 30 to 32 year return period. Peak History of small failures in this slope. pressures. Progressive weakening on flat slopes below toe of cut, 3 m max at toe na Fugro (1999d)
- Slide B pressures along preferential component movement. remained relatively intact.
hourly rate = 62 mm/hr. with time of old cut slope also a of cut.
flow paths).
contributing factor.
No obstruction or channelisation. Minor deposition
Slide during 10 hours of heavy rainfall Rainfall (infiltration into soil, Instability due to reduced strength of Slide of debris (compound - non-circular). Debris travelled over surface, flattening vegetation, local Initial construction in 1976. Shallow cut back in
Lai Cho Road, on 34 deg slope (lateral margins), most deposited at
HK17 Cut (364 mm), max hourly rainfall of 128.5 mm. Previous sliding in vicinity of this failure. possible development of saturated soil and perched groundwater Shallow. Sliding partly along colluvium / erosion of surface slope and local deposition of debris on na 1988 and installation of surface drains. No Halcrow (1999a)
Kwai Chung top of cut slope in rock, probably stopped by cut-off
1 in 50 yr event for rolling 4 hr rain. perched water in colluvium). in colluvium. saprolitic soil interface. margins. washout effects.
trench.

July & Aug 1994 generally wet. Heavy rain on


Small bus caused minor obstruction. No channelisation. Slide on 7/8/94 one of series of three slides. Two
Castle Peak 22/7/94 (317 mm) and 23/7/94 (43 mm). Rainfall (rising groundwater Instability due to rising groundwater Debris moved swiftly down the slope. Bus hit and Chan et al (1996),
Previous slide in 1982 and further slides in July Rotational slide in saprolitic soil, large Depostion mostly on flat slope at toe of cut (4 m max) with earlier failures on 23/7/94. One man killed on 7/8/94
HK18 Road (Milestone Cut Heavy rain on 6/8/94 (287 mm) and 7/8/94 levels following prolonged levels. Brittleness probably from tumbled down embankment. 5 to 6 subsequent minor na Franks (1995),
1994 (one month prior to this failure). number of defects in weathered rock. some extending down slope below road (thinly deposited when slide hit light bus and pushed it off the road.
14.5) very heavy (peak hourly 42 mm/hr). Little rainfall). internal brittleness of slide. slips over next 1 to 2 hours. Pun and Yeo (1995)
on flatter portion of steep slope) Original cut slope formed before 1954.
rain 1 hr prior to 23/7/94 slide.

1971 - number of slips in cut face and


No channelisation and limited obstructions given
movement above cut. Decfect controlled compound slide in Suspect increase in mobility due to presence
Rainfall (also contribution Instability due to loss of strength on Catastrophic failure at 9:00 pm on 18/6/72. Suspect size of failure. Limited deposition in failure scar Hungr GR (1998)
Heavy rainfall for 3 days prior to slide. 15 to 18/6/72 - large settlement above cut (> saprolite. Buttress effect at toe. 10 (DAN of surface water below toe of cut (particularly
HK19 Po Shan Road Cut from excavation at toe of saturation, perched water tables and rapid movement. Limited disturbance of material at rear (except at front), mostly in flat area at toe of cut Vail (1972)
Prolonged rainfall for approx 40 days. 4.5 m) increasing with time up to failure and Significant translational component of model). on flat cut surface). Material at rear of
natural slope). strain weakening due to movement. of deposited mass. (up to 12.5 m depth) reducing towards toe of Irfan and Woods (1998)
movement at toe of cut. Several slips rupture surface. slipped mass remained relatively intact.
slipped mass.
preceeded the large slide.

No channelisation, no obstruction. Most of debris


Rainfall (infiltration and Instability due to development of Defect controlled translational slide along Trees at toe of cut slope fell immediately prior to failure. Defect controlled landslide in granitic saprolite
deposited in base of slide scar (29 deg) and platform
Kau Wa Keng Heavy rainfall on morning of failure, development of perched perched water table on persistent slightly undulating discontinuity in Landslide was sudden and debris ran out rapidly. soils. Cut formed between 1954 and 1959.
HK20 Cut na at toe of cut (1 to 2 m depth). Some material na Halcrow (1998c)
Upper Village 280 mm in 6 hours, peak 129 mm/hr. groundwater table on discontinuity (discontinuity exposed in saprolite. Subvertical joints acted as Second, smaller slide occurred some 1.5 hours after Suggests possible weakening and opening of
continued down slope below platform and flowed
discontinuities). face of excavation). release surfaces. initial slide. joints in period up to slide. 1 fatality.
down to Castle Peak Road.

Muddy water observed out of weep holes for 2 Dislodgement of some masonry stones then wall "burst"
Heavy rainfall for 2 days preceeding slide Service leakage (saturation Compound (possibly rotational behind GEO (1994),
days prior to failure. 2.5 hours prior to slide Masonry block retaining wall unable to out at about mid height. Collapsed instantaneously. No obstruction of channelisation. Deposition on Collapse of retained cut slope (built before
HK21 Kwun Lung Lau Cut - R (1 in 28 year event). Only 29 mm rain in of supported soil, possible retaining wall) failure in retained saprolitic na Morgenstern (1994,
greater rate of seepage through weepholes and support saturated saprolite. Upper portion of failed wall remained relatively intact at flat slope, maximum 6 m depth at original toe of wall. 1904). 5 people killed.
10 hours up to failure. perched water). soil slope. 2000)
joints. toe of failed mass.
Appendix B Page B11

Table B1.3: Landslide case studies from Hong Kong (Sheet 3 of 4)


Landslide Geometry Width (m)
Slope Travel
Date of Slope Initial Slide Travel Obstr- Confined Slope Slide Slide Failure Ratio Volume Material Description
No. Site Failure Length, Alpha 1 Alpha 2 Alpha 3 Alpha 4 Distance 3 Initial Geology Hydro-geological Features Morphological Features
Failure Type Classification Classification ucted (Yes/No) Height, Height, Length, L L (m) Depth, D H/L (m ) Runout (Slide Source)
Geometry
Hs (m) toe (deg.) (deg.) (deg.) (deg.) Angle ( o) Slide
H (m) (m) (m)
Deep fill (up to 25 m) at slope of 34
Flow slide Loose fill - end dumped decomposed granite Significant stream course at failure location prior to
degrees over natural ground. Slope
HK22 Sau Mau Ping 25/8/76 Fill 1 Compound - significant flow slide Yes No 32 32 80 50 2 to 4 34 30 0 na 21.8 0.40 2500 42 42 Granite (SM - silty sand ?). 10 years old. 75 to 90% filling. Estimated depth of wetting from rainfall infiltration
and downslope entirely of loose fill.
translational component SMDD. of 2 to 6 m.
Vegetated.

Flow slide, compound - Fill placed in old valley. Maximum depth


flow slide / debris Fill - wet silty sand (SM) with pieces of brick Undistubed fill remaining post slide was wet.
HK23 Kennedy Road 8/5/92 Fill 1 significant translational No No 31.1 26.6 44 28 2 35 45 0 24 31.2 0.60 500 9 to 13 6 to 12 Granite. approx. 5 m at slope of 42.5 to 55
flow and stone. 50 years old. Infiltration from rainfall caused wetting.
component degrees.

flow slide / confined


Flow slide (part)
debris flow (for part Infiltration from rainfall and also from prolonged leakage of
Sha Tin Heights Compound - significant 34 Fill - loose silty coarse sand (underlain by Steep (34 up to 49 degrees) shallow fill
HK24 4/6/97 Fill 1 that flowed) debris Yes Yes (for flow) 47.5 39.5 86.5 70 1.5 27.5 0 27.5 24.5 0.46 150 15 9.5 Granite drainage pipes. Likely perched water table between fill
Road - Slide A translational basal (25 to 50) colluvium) slope on 25 to 35 degree natural slope.
slide for intact and natural.
component
portions.

3
Kau Wa Keng Flow slide Located along line of ephemeral stream. Infiltration of 1000 m of granitic fill (unretained) dumped
Fill - loose, moist to wet clayey silty gravelly
HK25 San Tseun - 4/6/97 Fill 1 Compound (?), possibly flow slide No No 20 22.5 52 30 3 to 4 36 to 40 32 0 19.5 23.4 0.43 500 11 11 to 12 Granite rainfall and water from stream flow into fill, with possible on natural slope. Near level fill platform
sand (decomposed granite). End tipped fill.
Slide D rotational, retrogressive development of perched water table in fill. and steep side slopes (36 degrees).

Infiltration of water into fill and leakage from stormwater


Chung Shan
Flow slide system above slope. Buried retaining walls potentially
Terrace, Lai 30 Fill - sandy silt to clayey sandy silt with gravel. Steep loose fill slope (37 degrees). Buried
HK26 4/6/97 Fill 1 Compound - significant flow slide No No 23.5 23 73 56 3 37 30 2 17.5 0.32 450 18 20 to 13 Granite dammed the flow of water resulting in elevated perched
King translational component
(21 to 37) Avg 75% SMDD indicating loose condition.
water table. Seepage observed from base of main scarp
retaining walls in fill.
Hill Rd and base of masonry wall.

Fill - clayey silty sand (loose to med dense) and Herringbone drainage on fill surface, cut off drain at top of
Compound - significant debris slide / debris 30 clayey sandy silt (soft to firm). 86 to 94% SMDD slope. Landslide located immediately below blocked Benched fill slope. Failure within upper
HK27 Tuen Mun Road 2/7/97 Fill 2 No No 11.5 11.5 22.2 7.7 1.2 0 - 34 27.4 0.52 200 25 25 to 28 Granite
translational component flow (50 at toe) for general fill. Probes - uniformly low density to drainage channel. Infiltration from rainfall and run-on from bench at slope of 34 degrees.
5 m depth. upslope. No seepage observed in main scarp.

Fill - silty and gravelly sand (SM), end


Flow slide Shallow illegally dumped fill at slope of
Au Tau Village dumped, loose, 70% SMDD with pieces of Infiltration from rainfall and run-on from upslope.
HK28 9/6/98 Fill 3 Translational (on flow slide No No >80 25.8 57.5 43.5 2 35 23 16.5 23 24.2 0.45 170 15 24 Volcanic - tuff 35 degree. Dumped on natural slope
Road building rubble. Underlain by colluvium (firm Leaking drainage pipe.
interface), retrogressive ? of 23 degrees.
to stiff clay).

No surface drainage above retaining wall. Run off over- Retained fill with relatively flat slope
Compound or Rotational, Fill - soil with building debris, loose, retained.
HK29 Baguio 8/5/92 Fill - R 1 confined debris flow Yes Yes 109 109 227 207 7.5 80 31.4 0 16 25.6 0.48 3800 35 11 to 15 na topped retaining wall. Infiltration from rainfall. Possible above wall. Natural gully below retaining
retrogressive Traced back to before 1924.
perched groundwater table within fill. wall.
Surface water from upslope flowing onto concrete fill Series of fill terraces supported by
Kau Wa Keng Fill - loose to very loose, wet, silty gravelly
Rotational, possible flow slide / debris 90 (ret platform. Springs flowing out of exposed rock face above retaining walls on 33 deg slope. Failed
HK30 San Tseun - 4/6/97 Fill - R 1 Yes No 80 10 21 17.5 2.5 33 0 33 25.5 0.48 60 7 6 to 9 Granite sand with cobbles, boulders, brick and
flow slide flow wall) fill. Saturation of fill and possible development of perched retaining wall had concrete covering in
Slide C concrete. Numerous voids.
groundwater. poor condition.
No drainage at top of slope. No drainage behind retaining
Tao Fung Shan Fill - firm to soft gravelly clayey sandy silt,
40 36 wall, some weep holes in face. Substantial run-off from Series of retaining walls on average
HK31 Cemetery 2/7/97 Fill - R 1 Compound - non circular debris flow No No 32.5 21.5 37.5 25.5 2.6 0 36 29.8 0.57 400 27 20 Granite overlying residual soil (firm gravelly clayey
(28 to 90) (34 to 46) up-slope observed, likely to have ponded behind retaining slope of 40 degrees.
- Slide B sandy silt).
wall and infiltrated retained fill.

7 to 12 (in
Possibly affected by discharge from upslope (erosion pipes
Slide of debris, gully), up
Sha Tin Heights 37 16 Colluvium - silty fine to coarse sand. Minor observed in scarp). Infiltration from discharge and rainfall. Steep natural slope (up to 42 degrees)
HK32 4/6/97 Natural 1 translational on confined debris flow No Yes 44 33.5 105 86 1.5 0 na 17.7 0.32 170 12 to 20 m on Granite
Road - Slide B (32 to 42) (13 to 22) filling. Perched water table developed in colluvium (above above flatter slopes of 13 to 22 degrees.
interface, retrogressive flatter
saprolite). Erosion pipes possibly from upslope discharge.
slopes

Groundwater at depth, no evidence of seepage. Possibly Detachment scar located within a locally
Wonderland Compound - significant
debris flow, lower Residual soils - completely decomposed run-on from upslope, report of flooding in development at steepened (40 degree slope)
HK33 Villas, 5/8/97 Natural 1 translational component on Yes Partly 107 104 190 177 1.0 to 1.5 40 34 0 na 28.7 0.55 80 8 6 to 8 Granite
part confined granite. slope crest. Rainfall infiltration and development of topographical depression at the head
Kwai Chung interface
perched groundwater table in residual soil. of a poorly developed drainage line.

No surface drainage at top of slope. Depression at top of


Compound - translational Saprolitic soils - completely decomposed Located within natural depression at head
Ka Tin Court, debris flow, partly crest allowed ponding of water. Infiltration promoted by
HK34 2/7/97 Natural 1 basal component (defect Yes Yes 26.6 23.5 43 32.6 2.5 52 37.5 0 na 28.7 0.55 150 14 8.5 Granite granite. Overlain by shallow fill (loose, end of gully. Failure in over-steepened slope
Shatin confined shallow loose fill on crest and upper slope. Groundwater
control ?) dumped silty fine sand). of 52 degrees.
at depth.

Saprolite - sandy silty clay (CI/CH) with Infiltration and seepage resulting in high perched water Natural slope at average angle of 31
Compound - non circular weak CH clay seams. Overlain by fill (loose table in main slide area, exaserbated by run-on from road degrees. Cut to fill at crest of failure
HK35 Shum Wan 13/8/95 Natural 1 debris flow No No 71 71 219 154 12.5 31 26.5 0 na 18.0 0.32 26000 50 60 to 90 Volcanics - Tuff
(defect control) sandy silty clay CL/ML, 71 - 81% SMDD) at crest following small fill failure. High groundwater table (Nam Long Shan Road, fill 5 m deep).
and colluvium (silt/clay of 1 m depth). in lower planar slide area. Slope densely vegetated.

Natural slope at average of 34 degrees.


Compound - some Saprolite - medium dense silty coarse sand
Ma On Shan debris flow, upper Concentration of surface and sub-surface water Minor cutting and filling on lower slopes
HK36 2- 3/7/97 Natural 2 defect control, No Yes, partly > 45 42.4 106.5 47.5 2.5 to 5 33.8 5 - na 21.7 0.40 3000 12 to 30 12 Granite with gravel. Persistent planar discontinuity in
Road part confined flow within gully. within gully. Failure located within gully
retrogressive saprolitic soils.
with previous history of landsliding.

Initial slide area on steep mountain slope


Colluvium - Cobbles and boulders in a matrix
25 to 30 50 to 95 Ephemeral stream. Flows strongly during and just after (50 degrees). Narrow V-shaped gully to
Slide of debris, 37 26.7 12.9 Granite, Sedimentary (15 to 25%) of gravel to clay size. Less
HK37 Tsing Shan 11/9/90 Natural 3 Compound
confined debris flow No Yes 480 379 875 715 4 na 23.4 0.43 13000 (and (depositi and Volcanics boulders and more finer material in lower
rainfall with some low flow (springs). Infiltration and Ch 500 m (avg 27 to 37 deg). Broad fan
(28 to 50) (16 to 40) (6 to 17) possible saturation of colluvium. area on lower hillslopes below Ch 500 m
gully) on zone) valley channel.
(17 reducing to 6 degrees).

Sedimentary and Natural slope (densely vegetated) with


Saprolite (completely decomposed tuff) and Infiltration through open jointed rock and development of
Outward Bound Compound - non circular debris flow, partly Volcanics - tuffaceous steep upper slope (44 degrees) at head
HK38 9/6/98 Natural 3 No Partly 60 32 63 48 3 to 5 44 33.5 19 na 26.9 0.51 900 21 16 to 8 residual soil - stiff clayey sandy silt. Overlain water pressures in joints and/or elevated water pressures
School (some defect control ?) confined / debris slide mudstone, siltstone and of gully. Slopes in gully below slide
by shallow layer of colluvium. in soil mass. Erosion pipes promoted sub-surface flow.
breccia. decrease from 33.5 to 19 degrees.

Natural slope of 14 to 33 degrees.


Slide of debris,
Steeper slope (30 degrees) in failure
Lantau Island - 4 and Compound - significant debris flow, partly 18.3 No information. Likely infiltration from rainfall
HK39 Natural 3 No Yes, partly >100 84 218 181 3 33 - na 21.1 0.39 1000 10 to 15 10 to 15 Volcanics Colluvium - 30% cobbles. area (colluvial accumulation). Debris
Slide C1 5/11/93 translational component on confined (14 to 23) resulting in saturation of colluvium.
flow within broad and not well
interface
formed gully.

Residual soil - loose to med dense gravelly No drainage at top of slope. Deep groundwater. Erosion
Tao Fung Shan 52 (two Natural slope at average angle of 40
Natural 6 clayey silty sand. Overlain by fill (tipped, loose pipes (in fill) exposed in backscarp. Spring approx. 30 m
HK40 Cemetery 2/7/97 1 Translational debris flow No No 63 60 104 89 3 to 3.5 50 40 40 30.0 0.58 900 backscar 50 to 65 Granite degrees. Steep slope (50 degrees)
/ Fill (11 to 0) to very loose, gravelly silty clayey sand of 1 to 2 below scarp. Infiltration from rainfall, possibly
- Slide A ps) in shallow fill on top of natural slope.
m thickness) with topsoil layer below fill. exacerbated by presence of fill.

Infiltration of water into slope. Surface depression at top of


Compound - translational Residual soil - loose to med dense sand
Glorious Praise Natural slope ponded water which overflowed onto slope. Erosion Average slope of 30 degrees. Shallow
HK41 8/5/97 1 basal component on debris flow No No 8.7 9 17.2 12 1.5 32.5 32.5 1 0 (?) 27.6 0.52 20 4.9 5 to 6 Granite overlain by fill (loose sandy silt / silty sand,
Christian Centre / Fill holes observed in fill in scarp (no seepage). Possible fill on relatively flat ground at top of slope.
interface 1 m thick).
perched water table between residual and weathered rock.

Infiltration of rain water into slope. Ponded water upslope of Natural slope (25 to 50 degrees) with
Residual soil (silty sand) overlying saprolitic
Natural scarp. Increase in groundwater level within slope. Seepage series of small terraces down the slope
HK42 Lido Beach 2/7/97 4 Compound - non circular debris flow No No 48 32 83 60.5 3 52 to 25 17 - na 21.1 0.39 750 24 14 to 27 Granite soils (silty sand). Shallow layer of fill (silty
/ Fill from base of rupture surface indicative of high groundwater (minor cuts and fills). Modified natural
fine sand).
table. slope.

Slide of debris, 8 to 10 Infiltration from rainfall and run-on from upslope. Flow in Natural gully - slopes
Compound - significant debris flow, partly 20 9.6 (gully), Sedimentary - Colluvium - silty sand with large content of topographic depression possible during heavy rain. decreasing gradually from 31 degrees
HK43 Lui Pok School 26/9/93 Natural 1 Yes Yes, partly 180 77 220 186 2 31 na 19.3 0.35 450 29 Sandstone gravel and cobbles. Seepage observed in scarp from erosion pipes and at scarp area to 10 degrees above
translational component on confined (29 to 15) (0 to 10) 34
interface terminal. bedding discontinuities. school.
Appendix B Page B12

Table B1.3: Landslide case studies from Hong Kong (Sheet 4 of 4)


Slope Post-Failure Characteristics Velocity
No. Site Rainfall Pre-Failure Observations Trigger Mechanism Initial Slide Description Post-Failure Observations Comments References
Type (Obstructions, deposition, channelisation) (m/s)

Hungr GR (1998)
Instability due to loss of strength on Flow slide in loose fill. Compound with
Heavy rain on 24 - 25/11/76, 15 to Some erosion and small periperal slides during Rainfall (infiltration into outer Upper part of slope moved as a whole. No water Obstructed by building. Deposition only on flat 10 to 15 DAN frictional model, phiB = 20. HKIE (1998)
HK22 Sau Mau Ping Fill saturation. Likely collapse of loose fill significant translational component.
35 mm/hr. 1972 heavy rainfall event. zones of loose fill). observed to drain from slipped debris. slope (none on 30 deg slope), max depth 1.8 m. (DAN) 18 people killed Morgenstern (2000)
structure resulting in static liquefaction. Shallow, entirely within loose fill.
Gov't HK (1977)
Instability due to loss of strength on Slight obstruction by car. Depostion on steep slope
163.5 mm in 2 hours leading up to slide Rainfall (infiltration into outer Flow slide in loose fill. Compound with HKIE (1998),
HK23 Kennedy Road Fill na saturation. Likely collapse of loose fill Debris wet. (< 1m depth) possibly due to batters in cut slope. 5 to 10 (est) Slide preceeded by collapse of "big" tree.
(1 in 20 yr event) zones of loose fill). significant translational component. Chan et al (1996)
structure resulting in static liquefaction. Greatest deposition on flat slope below cut.

Possible flow slide in loose fill. Flow portion of slide mass was confined. Some
Heavy rainfall on 4/6/97 - 296 mm prior to Instability due to loss of strength on Additional mobility from surface water flowing
Sha Tin Heights Compound with significant translational Part of slide flowed down confined travel path. Part obstruction from heavily vegetated gully. Stripped
HK24 Fill slide, with max hourly 81.5 mm/hr. 4 hr na Rainfall (infiltration into fill). saturation. Likely collapse of loose fill na in gully. Query flow slide classification, possibly Halcrow (1999a)
Road - Slide A component along interface between loose remained intact and moved only short distance. vegetation in gully. Most of material deposited on flat
incident 1 in 22 year event. structure resulting in static liquefaction. dilatant failure of fill on steep hillslope.
fill and natural. slopes at toe (No deposition on 25 to 35 deg slopes).

Instability due to loss of strength on


Kau Wa Keng Rainfall (infiltration and Possibly two phases of failure - initial failure in steep Fill placed about 1961. Part of fill located on relict
24 hr - 263 mm, 12 hr - 252.5 mm, max saturation and perched groundwater in Flow slide in loose fill. Compound (?), No obstruction or confinement of flow. Deposition
HK25 San Tseun - Fill Initial small slide at front of fill body. possible development of slope followed by washout failure caused by flow in na landslide debris. Possible extension of slide mass due Halcrow (1999a)
intensity 128.5 mm/hr (1 in 40 yr). fill. Likely collapse of loose fill structure possibly rotational, retrogressive on 32 deg slope and flatter.
Slide D perched water in fill). ephereral stream. 4 m deep channel cut in run-out path. to washout (initial travel angle possibly 30 deg).
resulting in static liquefaction.

Chung Shan Rainfall (infiltration and


Heavy rainfall in 2 hours preceeding slide, Instability due to loss of strength on Landslide sudden and fast moving, liquefaction likely. Some scouring of downslope fill. No obstruction or
Terrace, Lai saturation of loose fill, Flow slide in loose fill. Compound with 1949 - terraced natural slope evident. 1949 to
HK26 Fill peak intensity of 128.5 mm/hr (1 in 41 na saturation. Likely collapse of loose fill Following slide, surface water flow eroded some material channelisation. Deposition predominantly on flat na Halcrow (1999b)
King possible perched water significant translational component. 1963 - placement of filling.
year event). structure resulting in static liquefaction. and resulted in washout for considerable distance. slopes at toe.
Hill Rd table).

Not liquefaction of fill. Failure due to wetting Dilatant failure through soil mass. Failure and travel completely on fill slope. Fill slopes
24 hr - 186 mm, 12 hr - 178 mm, 1 in 6 year
of the fill from infiltration of rainfall and run- Compound with significant translational Photos indicate only partial break up of sliding mass, thus constructed 1972 to 1975. Cut at toe of fill slope in
HK27 Tuen Mun Road Fill for rolling 15 min rainfall intensity. 665 mm na Rainfall Majority of debris remained in source area. na Halcrow (1999a)
on due to blocked drain. Small cut at toe component. Shallow, not particularly debris slide / debris flow. 1996. Slope vegetated with grass. Compare with cut
in 31 days prior to slide.
probably ontributed to instability. mobile. slopes.

No obstructions or channelisation, width almost


Most severe event since mid 1980s for 2 to 24 Instability due to loss of strength on Flow slide in loose fill. Shallow Crossfall of road altered in months preceeding
Au Tau Village doubled, therefore flow spread. Deposition - Less
HK28 Fill hour rainfall. 12 hr rainfall 323 mm (18 year na Rainfall (infiltration). saturation. Likely collapse of loose fill translational slide on interface between fill Some secondary outwash followed initial debris run-out. na slide resulting in ponding above fill slope Fugro (1999b)
Road than 0.5 m thick on 23 degree slope, > 1m thick
event), intensity 30 to 60 mm/hr at peak. structure resulting in static liquefaction. and natural, retrogressive (?) and run-on onto fill slope.
on 16.5 degree slope at distal end slide.

Heavy rain on day of failure 350 mm. Initial wall failure due to either erosion at Possibly dilatant failure in retained fill Failure occurred as a series of retrogressive failures of Building obstructed flow at toe. Confined within
3 3 Fluidisation of debris by water flowing in natural
HK29 Baguio Fill - R Short periods of very heavy rainfall, na Rainfall. toe (from over-topping) and/or perched slope. Compound or rotational, 200 to 1000 m over a period of 5 to 6 hours. Flow in natural gully. Accumulation (1500 m ) in gully. na Chan et al (1996)
gully.
1 in 60 yr for 5 min rainfall prior to slide. water table in retained fill. retrogressive. natural gully contributed to mobility of the flow. Deposition on flat slopes on and below Victoria Rd .

Kau Wa Keng Rainfall (infiltration and Instability due to reduced strength of Rotational failure of retained fill. Possible Obstructed by house. No channelisation. Most of In 1996 no seepage nor distress of slope were
24 hr - 263 mm, 12 hr - 252.5 mm, max Single rapid movement followed by secondary washout
HK30 San Tseun - Fill - R na development of perched saturated soil and perched groundwater development of flow slide once rataining deposition on flatter slope (max 2 m depth), minimal na observed. No comment on extension of debris due to Halcrow (1999a)
intensity 128.5 mm/hr (1 in 40 yr). scarp 6 m back into slope and 2m deep in runout path.
Slide C water above saprolite). above weathered rock. wall had failed. on terraced slope (avg 33 degrees). washout. Possible flow slide once retaining wall failed.

3
Tao Fung Shan 971 mm in 31 days prior to slide. Heavy Instability associated with ingress of Likely rapid movement once retaining walls failed. No channelisation. 150 m debris picked up by slide. 1986 - 1990 retaining walls constructed. Some
1996 - Retaining walls 1 and 2 reported as Rainfall (infiltration into
HK31 Cemetery Fill - R rain prior to slide, peak 1 hour intensity water into fill. Water table development Compound - non-circular Possible liquefaction of fill. Secondary outwash (water Minimal deposition on steep slopes (34 to 46 deg), na secondary outwash, 1 m deep erosion through residual Halcrow (1999b)
being unstable. Vertical cracking evident. retained fill).
- Slide B of 124 mm (1 in 32 year event) due to absense of effective drainage. mainly from run-on from upslope) of fines significant. most deposited on flat slope at toe. soil and channelised outwash 30 m further downslope.

Additional mobility from surface water flowing in gully.


Translational slide of debris along 3
Instability due to reduced strength of May have occurred in 3 stages. Initial failure (130 m ) Confined in gully. Minor obstruction. Deposition on Failure associated with ongoing processes of gully
Heavy rainfall on 4/6/97 - 240 mm prior to interface between colluvium and residual
Sha Tin Heights saturated colluvium and perched flowed and 2 subsequent intact translational movements slopes flatter than 13 degress in gully. Below gully development. Greater mobility than Slide A possibly
HK32 Natural slide, with max hourly 81.5 mm/hr. 2 hr na Rainfall (infiltration). soil. Retrogressive - first stage of 130 m3 na Halcrow (1999a)
Road - Slide B groundwater on colluvium / residual of short distance. Travel distance and velocity of first approx 0.1 m thick deposition on flat slope. Possibly due to greater catchment area and failure during more
incident 1 in 9 yrs. followed by two smaller intact
soil interface. stage probably influenced by surface water flow. affected by washout to some degree. intense period of rainfall. Washout likely to have
movements.
extended slide mass.

Deposition on outer margins of flow (minimal)


Instability due to reduced strength of Initial failure at 8:00 am, fully developed at 10:00 am.
Wonderland 8/5/97 - 160 to 200 mm starting 6 hrs prior Relict landslides within nearby drainage lines. Rainfall (infiltration and most in pipeline and road below (Not on 33 to 35 Travel angle = 34 degrees to drainage pipe, 29
saturated colluvium and perched Compound with significant translational Two phases of movement. Slide mass over-rode channel
HK33 Villas, Natural to failure, intensity peaking at time of slide No evidence of previous instability within this development of a perched deg slopes). Lower part of flow confined in drainage na degrees to toe of slope. Possible washout Halcrow (1999a)
groundwater on residual / weathered component along weathered rock surface. stripping vegetation and some surface soils, then entered
Kwai Chung (I hr rainfall 1 in 12 yr event). drainage line. water table). line. Obstruction was constriction of flow into effect.
rock interface. a drainage line and washed down onto road below.
drainage pipe.

Partly confined in short gully section. Obstructed by Surface fill on slide area end tipped 1978 to 1980. Cut
Heavy rainfall. 31 day - 1108 mm. 12 hr - Compound failure (translational with steep Slide debris stripped surface soils from downslope path in
Ka Tin Court, Instability due to loss of strength on river at toe of slide. Most of debris deposited at toe at toe of slope 1980 to 1984. Water flow in gully may
HK34 Natural 255.5 mm. 1 in 20 to 1 in 45 yr return na Rainfall rotational backscarp) through saprolitic gully and partly blocked river. Washout of debris blocked na Halcrow (1999d)
Shatin saturation. Exaserbated by surficial fill. of slope within water (max 4 m depth). Stripped have increased the mobility of the slide. Densely
period for 1 to 4 hr events. soil mass, possibly some defect control. draining cascade.
surface soils in gully on 37 deg slope. vegetated slope.

Initial small fill failure due to infiltration,


Main slide - not rapid but continuous, significant degree Surface water probably contributed to mobility. Hungr GR (1998)
Rainfall (infiltration into fill which resulted in surface water flow over Main slide - partly defect controlled No channelisation or obstruction. Deposition - planar slide
Heavy and prolonged rainfall. 31 day - 1 of break up. The lower planar slide remained relatively 19 (DAN) Significant control of discontinuities in main failure and Irfan (1997)
Observation of muddy water on hillslope days caused initial small fill failure. slope. This (with infiltration) resulted in high compound slide in saprolitic soil. Lower material on flat slope (2 to 3 m depth), main slide
HK35 Shum Wan Natural in 75 yr event. 4 hour - 159 mm (peak intact. Followed by several retrogressive slides which maybe bit also planar failure. 2 people killed. DAN model - OK Irfan and Woods (1998)
prior to failure. Main failure - Surface flow, perched water table that caused main slide. slide - translational planar slide on deposited mostly on hillslope (27 degrees) and in base of
48 mm/hr). took out Nam Long Shan Road. Surface water likely to high. with phiB = 20 deg, Voellmy model better GEO (1996b)
infiltration and perched water). Planar slide caused by weight of debris from kaolinite seams in PW tuff. main scarp (3 to 5 m depth).
have contributed to mobility. representation. Knill (1996b)
main slide.

Heavy rainfall. 1 and 2 day events very Slide through saprolitic soil mass, some Initial failure possibly on lower cut and fill slopes due to Gully with history of landsliding. Densely
Instability of main failure possibly No obstruction. Mid to upper part confined.
Ma On Shan severe (1 in 1000 yr return period). Max. defect control (slide partly along surface water concentration in gully and subsurface water vegetated slopes and grassed lower
HK36 Natural na Rainfall caused by smaller failures in cut and fill Deposition on lower slopes (minimal on steeper na Halcrow (1999c)
Road hourly = 94.5 mm/hr and daily = 796 persistent planar discontinuity). flow. Triggered further regressive landsliding. Possibly flat slopes. Considerable erosion of debris
slopes at toe. Some defect control. slopes of 30 to 55 degrees) of 2 to 2.5 m depth.
mm/day. Compound and retrogressive. fast moving. and washout.

Initial slide triggered by rainfall. Main Loud noise heard for period of 5 to 30 mins. Channelised No obstruction. Channelisation to Ch 500 m. 16.5 to 11.5 (15 to
Heavy rainfall, but not exceptional. 138 Slide of debris in colluvium mantling Channelised debris flow in colluvium on steep
slide triggered by flow of initial slide debris flow to Ch 500m which accumulated colluvium in Minor deposition within gully. Most of deposition 25 frict. model, < 20 Hungr GR (1998),
HK37 Tsing Shan Natural mm in 5 hrs (< 1in 2.5 yrs event for all na Rainfall slope. Possible signs of being 3 hillside. Interesting that triggered by low return
on to it. Possible liquefaction in coarse gully (approx 9000 m ). Deposition beyond Ch 500 m on on 17 decreasing to 6 degrees slopes, and where using Voellmy King (1996)
storms < 5 hr duration). retrogressive. period rainfall.
colluvium given open void structure. slopes of 17 decreasing to 6 degrees. width of flow widened to 50 to 95 m. model)

Instability due to adverse water table Relatively intact rafts of material remained within Minor obstruction by dense vegetation. Lower
Very heavy rainfall over 12 to 24 hour Rainfall (infiltration and Compound (non-circular) failure through
Outward Bound conditions. Slope already in meta-stable landslide scar, most material developed into debris flow. portion of slide partly confined in gully. Deposition Would expect surface water flow in gully to
HK38 Natural period prior to failure (1 in 30 to 70 year na development of perched residual and saprolitic soil mass. Possible na Fugro (1999c)
School condition. Possible adversely Initial slide followed by washout further down gully of in failure bowl and on all slopes downslope of failure have increased the mobility of the slide mass.
events) water). influence of discontinuities.
orientated faulting. sandy silt. (33.5 to 19 degree slopes).

Slide of debris in colluvium mantling Eroded channel to Ch. 150 m. Deposition from Large runout distance for suspected"gravitational"
Considered by GEO as a "gravitational" debris slide.
Lantau Island - Instability due to saturation and slope. Compound, with significant failure bowl to distal toe of debris on slopes of 11 @ ch. 115 (16 to debris slide (GEO). Classified as partly confined debris Hungr GR (1998),
HK39 Natural Severe rainfall for 2 days. na Rainfall Partly confined within hillslope topographic depression.
Slide C1 over-steepening of colluvium. translational component partly along 14 to 23 degrees. Most of finer debris deposited 8 from DAN) flow. DAN model, successful with phiB = 23 degrees, Wong et al (1996)
Eroded channel (< 1 m depth) to Ch 150 m.
interface with saprolite. to Ch 150 m, thereafter boulder trail. good velocity correlation.

3
Instability associated with fill surcharge, 300 m debris accumulated. Deposition on lower
Tao Fung Shan 970 mm in 31 days prior to slide. Heavy 1988 to 1996 - Loose fill placed at crest of
Natural saturation of fill from infiltration (both Translational - single event, base of slide Diverged into two debris trails. Outwash followed initial slopes at less then 11 degrees (No deposition on
HK40 Cemetery rain prior to slide, peak 1 hour intensity na Rainfall (infiltration). na slope. Densely vegetated slope. Some Halcrow (1999b)
/ Fill rain and run-on). Possible liquefaction in residual soil. slide (from surface runoff only ?). 40 degree slopes). No channelisation. Dense
- Slide A of 124 mm (1 in 32 year event) outwash.
and flow of loose fill. vegetation probably obstructed flow to some degree.

Not sure why low travel angle, because only small


Rainfall (infiltration and Compound slide in residual soil with Rapid slide. Flowed down 32.5 deg slope to broad No obstruction. No channelisation. Deposition
Glorious Praise Natural Heavy rainfall prior to slide - 90 mm/hr Instability due to saturation volume. Maybe some of secondary outwash included
HK41 na saturation, possible perched significant basal translational component channel. Flow down shallow slope of channel probably on 30 degree slope and on shallow slope in broad na Halcrow (1999b)
Christian Centre / Fill max hourly fall. (2 hr event 1 in 8 year). from rainfall. in calculation (several metres would make a significant
water). along interface with weathered rock. secondary outwash post initial slide. channel.
difference for this small slide).

Rainfall (rising groundwater High mobility possibly related presence of surface


Natural 31 day rainfall - 492.5 mm. 24 hr - 186 mm. following rainfall, 2.5 hr lag Instability due to loss of strength on Compound (non-circular) slide through No significant obstruction or channelisation. water on travel path and possibly internal brittleness in
HK42 Lido Beach Previous small slides in area in 1982. Two phases of movement. Quick moving. na Halcrow (1999b)
/ Fill 12 hr - 181.5 mm. between heavy rainfall and saturation and rising groundwater level. residual soil mass. Deposition on most of length of runout path. slope. Slope heavily vegetated. Developed 1954 to
failure). 1969. Filling in area of slip placed in 1980s.

Obstructed by building at toe. Partly confined in


Slide of debris in colluvium mantling
Instability, possibly due to development Likely to be fast moving. Entrained water in hillside hillside topographic depression. Erosion and Travel angle may be slightly over-estimated,
Heavy rainfall - 240 to 250 mm on slope. Compound with significant
HK43 Lui Pok School Natural na Rainfall of perched water and saturation of topographic depression likely to have increased the accumulation in topo depression. Deposition on na the runout length may include some post King (1997)
26/9/93. translational component along interface
colluvium. mobility of the slide. Partly confined. lateral margins on hill slopes of 10 to 20 degrees. failure washout.
with weathered bedrock.
Most material deposited on flat slopes at toe.
Appendix B Page B13

Table B1.4: Landslide case studies from Hong Kong (Ayotte and Hungr 1998)
Slope Landslide Geometry Travel Width (m)
Initial Slide Travel Confined Ratio Volume
No. Site Slope Type Failure Distance
Classification Classification (Yes/No) Height, Length, Length, Alpha 1 Alpha 2 Alpha 3 H/L (m3) Initial
Geometry
Ltoe (m) Angle (o) Runout
H (m) L (m) (deg.) (deg.) (deg.) Slide
TC - 6A/1, Tung
HK44 Natural 3 Slide of debris (?) Debris flow No 85 93 83 39 43 - 42.4 0.91 100 9 9
Chung
TC - 5A/10, 31 (35 to
HK45 Natural 3 Slide of debris (?) Debris flow No 50 73 60 44 - 34.4 0.68 300 12 25
Tung Chung 29)
Debris flow (confined
TC - 5A/13,
HK46 Natural 3 Slide of debris (?) along part of travel No 58.5 122 105 31 25 - 25.6 0.48 700 17 9
Tung Chung
path)
Lantau Island - 26 (28 to
HK47 Natural 3 Slide of debris (?) Debris flow No 89 180 160 30 - 26.3 0.49 400 16 7
Slide A6 23)
JK 419, New Debris flow, partly 38 (39 to
HK48 Natural 3 Slide of debris (?) No 157 191 175 39 - 39.4 0.82 1000 7 10
Territories confined. 32)
JK 515, New 26 (28 to
HK49 Natural 3 Slide of debris (?) Debris flow No 59 120 105 27 - 26.2 0.49 300 19 19
Territories 20)
Debris flow (confined
JK 529, New 36 (38 to
HK50 Natural 3 Slide of debris (?) along part of travel No 135 227 210 34 - 30.7 0.59 400 12 12
Territories 34)
path)
JK 410, New
HK51 Natural 3 Slide of debris (?) Debris flow No 63 100 85 28 33 - 32.2 0.63 400 10 15
Territories

HK52 Luk Keng Natural 3 Slide of debris (?) Debris flow No 57 150 120 30 24 0 20.8 0.38 170 10 6
Debris flow, partly
HK53 Pak Sha Wan Natural 3 Slide of debris (?) No 22 80 70 22 27 - 15.4 0.28 150 9 8
confined.
Debris flow, partly
HK54 Pat Sing Leng 1 Natural 3 Slide of debris (?) No 62 100 55 34 34 - 31.8 0.62 300 14 12
confined.

Cut / Quasi Compound, part


HK55 Fo Tan Station 1 Debris flow No 22.3 37.5 30 40 48 7 to 0 30.7 0.59 85 9 9
Natural ? defect-control

TC - 5A/2, Tung Channelised debris 25 (29 to


HK56 Natural 3 Slide of debris (?) Yes 129 312 275 32 13 22.5 0.41 1500 30 7
Chung flow 23)
Channelised debris 23 (36 to
HK57 Sha Tau Kok Natural 3 Slide of debris (?) Yes 331 1005 980 42 5 (6 to 2) 18.2 0.33 1500 27 10
flow 13)
Channelised debris
flow (poss. 23 (29 to 13 (16 to
HK58 Pat Sing Leng 2 Natural 3 Slide of debris (?) Yes 225 600 580 36 20.6 0.38 50 9 5
hyperconcentrated 19) 10)
stream flow)
Appendix B Page B14

Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles


Estimated Volume
Travel Classification Slope and Travel Geometry Width (m) Area (sq.m.)
(cu.m.)
Date of Initial Slide
No. Site / Stockpile Cross-section S/pile Travel
Failure Classification Height, H Length, L Length, Alpha 1 Alpha 2 Alpha 3 Alpha 4 Initial / Slide Deposit
Classification Geometry / Height, H s Distance H/L Initial Slide Flow Run-out
(m) (m) L toe (m) o (deg.) (deg.) (deg.) (deg.) Failure Area Area
Plan Shape (m) Angle ( )

National, South rotational ?, part flow slide / confined Confined, linear, 15000 to
SW1 3 Nov 1898 125 60 300 170 22.6 0.42 36 15 0 17 to 25 < 5000 40 5 to 10 - -
Wales flow slide debris flow (?) obstructed at toe 20000

Unconfined, linear,
Fforchaman, South
SW2 22 Feb 1935 rotational ? debris slide ? obstructed at toe by 100 75 290 180 19.0 0.34 36 27 10 27 40000 20000 90 100 - -
Wales river (?)

Abergorchi, South rotational, part flow debris slide / confined


SW3 4 Nov 1931 Confined, curved 162 70 740 640 12.3 0.22 36 8 - 18 100000 < 50000 90 20 to 30 - -
Wales slide ? debris flow

Craig-y-Duffryn, rotational, part flow Unconfined, linear,


SW4 16 Dec 1910 flow slide ? na 29 190 (?) 60 na na 36 15 - 15 + na na 90 120 - -
South Wales slide ? obstructed

Cilfynydd
100000 to
SW5 (Abercynon), South 5 Dec 1939 rotational, flow slide flow slide Unconfined, linear 130 46 590 530 12.4 0.22 36 15 (19 to 11) 4.5 19 100000 (?) 115 90 to 140 - -
130000
Wales

Nantewlaeth, South (rotational ?), part


SW6 1947-60 (?) flow slide Unconfined 40 27 170 120 13.2 0.24 36 8 - 8 < 20000 10000 40 50 to 20 - -
Wales flow slide

Maerdy, South Partly confined, 60000 to 20000 to


SW7 18 Nov 1911 rotational, flow slide flow slide
linear (obstructed ?)
105 60 440 320 13.4 0.24 36 11 - 0 (?) 90 100 to 50 - -
Wales 80000 30000

rotational (some
Aberfan - Tip 7, 12.8 (11.5 to
SW8 21 Oct 1966 retrogression), flow flow slide Unconfined, linear 177 71 690 580 14.4 0.26 36 12.8 8 to 8.5 150000 (?) 110000 80 100 to 220 - -
South Wales slide
14.8)

Aberfan - Tip 7, rotational, part flow 12.8 (11.5 to


SW9 Nov 1963 flow slide Unconfined, linear 106 68 390 315 15.2 0.27 35 12.8 - 10000 3000 75 80 to 40 - -
South Wales slide 14.8)

Partly confined,
Glenrhondda, (rotational ?), part 30000 to 15000 to
SW10 1943 flow slide linear, deflected at 108 52 360 280 16.7 0.30 36 11 (5 to 20) 0 (?) 17 65 70 to 20 - -
South Wales flow slide 40000 20000
toe.

Aberfan - Tip 4, rotational, part flow Partly confined, 12.2 (11 to 30000 to
SW11 21 Nov 1944 flow slide 184 85 600 450 17.0 0.31 31 (27 to 33) - 16 (9 to 18) 170000 115 140 to 35 - -
South Wales slide linear 18) 50000 (?)

Mynydd Corrwg (translational basal


Unconfined, 30000 to 20000 to
SW12 Fechan, South 17 Oct 1963 component ?), flow flow slide
obstructed (?)
100 40 320 235 17.3 0.31 36 17 - 17 to 30 100 110 - -
slide
50000 40000
Wales

Bedwelty, South (rotational ?), part 80000 to


SW13 12 Nov 1926 flow slide Unconfined 120 65 340 250 19.4 0.35 36 13 - 13 50000 100 110 to 50 - -
Wales flow slide 100000

Rhondda Main, Unconfined, linear, 2000 to


SW14 16 Feb 1928 rotational, flow slide flow slide 75 36 200 140 20.5 0.38 36 23 - 23 6000 30 15 to 30 - -
South Wales (obstructed ?) 4000

Pentre, South (rotational ?), flow 40000 to 20000 to


SW15 4 Feb 1909 flow slide Unconfined, linear 115 44 300 na 20.9 0.38 36 15 - 20 70 120 to 30 - -
Wales slide 60000 30000

Cefn Glas, South Partly confined, 40000 to


SW16 3 Jan 1925 flow slide flow slide 140 60 310 190 24.2 0.45 36 22 - 22 30000 90 50 - -
Wales linear, obstructed (?) 50000

rotational ?,
debris flow ? / flow
SW17 Parc, South Wales 1965 possibly out-burst, Unconfined 88 32 213 140 22.4 0.41 36 12 to 23 - 15 to 19 10000 10000 45 45 to 90 - -
slide ?
flow slide ?

flow slide ? / debris


rotational ?,
Fernhill, South flow (poss. Hyper- Unconfined,
SW18 4 Dec 1960 possibly out-burst, na 42 na na na na 36 15 0 (shallow) 15 (?) <10000 <10000 50 na - -
Wales concentrated deflected
flow slide ?
streamflow)
Brownie (FCL), Flow slide, Partly confined,
BC1 11/6/83 flow slide 420 235 900 590 25.0 0.47 38 na 15 25 110000 - 60 65 18612 -
British Columbia translational curved

Brownie (FCL), Flow slide, Partly confined,


BC3 21/9/84 flow slide 408 210 925 640 23.8 0.44 38 na 14 27 500000 - 170 250 47124 -
British Columbia translational linear
Appendix B Page B15

Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles (Sheet 2 of 6)

Date of
No. Site / Stockpile Hydro-geological Features Rainfall / Snowmelt Morphological Features Pre-Failure Deformation Behaviour Post-Failure Observations Comments References
Failure

Overtipping onto streambeds. Spring lines Relatively rapid. Temporarily stopped by Channelised runout of 5 to 10 m width confined
National, South 1 to 5 year antecedent event for 1
SW1 3 Nov 1898 associated with Birthdir coalseam. Slide confined na Indications of active landsliding. retaining wall that eventually gave way resulting to stream bed. Obstructed by retaining wall. Siddle et al (1996)
Wales to 25 days prior to slide.
to region of tip located over streambeds. in rapid debris flow. Most of material deposited at toe (?).

75 mm in 2 days prior to slide. Before Relatively slow moving, 0.4 m/min reducing to 0.1 Slide also in 1928. Both slides relatively slow
Fforchaman, South na na
SW2 22 Feb 1935 this only minimal rainfall. na m/min after some 6 hours. Decribed as "like sliding moving. Periodic tipping on this stockpile at Siddle et al (1996)
Wales
lava". Terminated at river at toe of slope (?) time of failure.
130 mm on day prior to slide and 160 - 170 mm Debris slide (poss. flow slide) that developed into debris
Abergorchi, South Tip covered parts of stream beds. Relatively slow moving debris flow (0.5 m/min)
SW3 4 Nov 1931 for 2 day total prior to slide. Prior to this virtually na na flow within narrow stream channel. Not clear if tipping Siddle et al (1996)
Wales Failure confined to stream bed areas. within narrow stream channel on 8 degree slope.
no rainfall for 23 days. active in area at time of failure.

Artesian pressures in underlying 150 mm in 5 days prior to slide. Tip possibly located on pre-existing Relatively rapid flow. Flowed down slope and Obstructed by existing canal and railway line. Possibly
Craig-y-Duffryn, argillaceous beds. Failure occurred 1 to 5 year antecedent event over 1 landslide. Shallow landsliding
SW4 16 Dec 1910 na into the canal and onto the railway line at the triggered by movement of pre-existing landslide following Siddle et al (1996)
South Wales during period of stormy weather. to 25 days prior to failure. observed on adjacent hill slopes base of slope (obstructed). period of heavy rainfall. Tip active at time of failure.
Tip covered two spring lines, one associated with Bishop et al (1969),
Limited rainfall (70 mm) in 9 days prior "ominous move" on tip on day before failure as Described as a flowing in a sheet of 90 to Suspect failure triggered by localised toe
Cilfynydd Daran-ddu coal seam and lower spring associated Upper part of tip founded on Bleasdale (1969),
to slide. Prior to this (Days 10 and 11), observed by coal pickers. Prior to failure a huge 140 m width. Rate of movement approx instability of that part of tip that had
SW5 (Abercynon), South 5 Dec 1939 with Boulder clay mantling slope. Runoff observed permeable scree. Lower toe Woodland (1969),
2 day rainfall of 90 mm (1 in 3 year) bulge developed in the lower half of the tip. 3 to 4.5 m/sec. Up to 12 m thick deposition on encroached onto Boulder clay foundation.
Wales to enter fissures (due to subsidence) in sandstone underlain by Boulder clay. H.M.S.O. (1967), Bishop (1973),
and 155 mm in 5 days. Survey suggests toe had bulged prior to failure. flatter slopes at toe. Tip active at time of failure.
above tip. Siddle et al (1996)
Nantewlaeth, South Small volume flow slide that ran-out on slopes of Tipped material included washery waste in
SW6 1947-60 (?) na na na na Siddle et al (1996)
Wales 8 degrees. Small flows in gullies as distal end. addition to normal overburben waste

Active movement of crest of tip in month prior to Possible pre-failure movements associated with movement
Tip overlies spring line. Active 80 mm on day prior to failure Run-out flowed down hill slopes and crossed
Maerdy, South Suggestion that lower portion of tip failure following heavy rains (crane tippers to be of pre-existing landslide following heavy rainfall.
SW7 18 Nov 1911 movement of tip associated Antecedent railfall equivalent to 5 to 20 colliery sidings. Terminated at Rhondda Fach Siddle et al (1996)
Wales located over pre-existing landslide. withdrawn from crest). Suspected similarity Foundation comprised thick deposits of colluvium and
with heavy rain. year event over 25 days prior to slide River where landslide debris dammed river.
leading up to reported failure. glacial till. Active tipping at time of failure.

Rapid transformation into flow slide that travelled quickly (rate


Approx. 6 m settlement behind crest in less than Active tipping at time of failure. Initial failure in Bishop et al (1969),
4.5 to 9 m/s) down slope. Dominant mechanism of
Tip covered stream and spring line. Pennant 12 hours prior to slide. Scarps behind crest toe region on pre-existing basal shear surface Bleasdale (1969),
Not unusual. 1 in 0.5 year event in Boulder clay at toe firm to stiff. movement as flow sliding on a liquefied basal component.
Aberfan - Tip 7, Sandstone highly permeable. Springs associated developing 3 - 4 months prior to failure, increasing followed by several retrogressive larger slide Woodland (1969),
SW8 21 Oct 1966 days preceeding slide (82 mm in 3 Suspect down-slope conditions Run-out spread as travelled down-slope and split into two
South Wales with Boulder clay mantling slope. Fluctuating pore in rate leading up to failure. Indication of previous components. Of estimated 150,000 cu.m. in H.M.S.O. (1967), Bishop (1973),
days). No rainfall on day prior to slide. wet and boggy at time of failure. separate components. Out-burst of ground water followed
pressure conditions at toe associated with rainfall. failure in toe area as well as back sapping. Toe initial failure approx. 110,000 cu.m. involved in Siddle et al (1996)
flow slide. Deposition on 12.8 deg slope (approx 1 m thick)
bulging. flow slide. Johnson (1980)
and at toe (up to 6 to 7 m deep).
Bishop et al (1969),
Tip covered stream and spring line. Foundation Active tipping at time of failure. Possibly
Bleasdale (1969),
Aberfan - Tip 7, groundwater levels and pore pressures at toe of
Approx 395 mm in month of Nov. Boulder clay at toe firm to stiff. Indication of previous failure in toe area as Approx 3000 cu.m. of estimated 10,000 cu.m. in triggered by heavy rainfall in Nov 1963 and
SW9 Nov 1963 slide suspected to have been significantly affected
Woodland (1969),
South Wales well as back sapping. failure developed into flow slide. localised instability in toe region. Wet and
H.M.S.O. (1967), Bishop (1973),
by rainfall. boggy down-slope conditions.
Siddle et al (1996)
Run-out flowed down hill slopes to Nant Selsig Active tipping at time of failure. Movement of
Glenrhondda, na na tip located on pre-existing landslide
SW10 1943 na River where it was deflected and flowed pre-existing existing landslide may have Siddle et al (1996)
South Wales
downstream. destabilised toe.
Bishop et al (1969),
Constructed over stream. Possible Bulk of failed material remained within bowl of
None for 2 days prior to slide, 160 mm Some sliding of tip prior to failure. Jan 1944 - one Active tipping at time of failure. Initial slide Bleasdale (1969),
Aberfan - Tip 4, perched water table above mudstone slide, only thin tongue travelled a long
SW11 21 Nov 1944 in 4 days prior to that (equivalent to na or two large movements of the tip occurred. Toe volume approx 170,000 cu.m., 40,000 cu.m. Woodland (1969),
South Wales seam. Large fluctuations in water distance (approx 40,000 of 170,000 cu.m.).
approx 5 to 10 year antecedent event). bulging for many months prior to failure. of this travelled a long distance. H.M.S.O. (1967), Bishop (1973),
level in Pennant Sandstone. Depth of deposition 11 to 2 m.
Siddle et al (1996)
Mynydd Corrwg 30 mm 3 days prior to failure. Overall,
Failure incorporated argillaceous Reference to displacement of natural ground at Flowed down to stream at toe of hillslope and Failure incorporated foundation of argillaceous
SW12 Fechan, South 17 Oct 1963 na significantly below 1 year antecedent rainfall for Siddle et al (1996)
bedrock foundation. toe prior to failure. blocked it (obstructed). bedrock.
Wales period of 25 days prior to slide.

None in 2 days prior to slide. Prior to Evidence of movement at toe. Movement of pre-existing
Tip straddled spring associated with
Bedwelty, South this, rainfall approx. between 1 to 5 Tip located on pre-existing "Boot" evident at toe of spoil heap adjacent to existing landslide may have destabilised toe. Majority of
SW13 12 Nov 1926 Birthdir Seam. Rainfall contributed to Halted by impact against wall at rear of school. Siddle et al (1996)
Wales year antecedent rainfall over Days "Pochin" landslide. area of slide. spoil heap affected by pre-existing landslide. Suspect
tip instability.
8 to 18. active tipping in this area at time of failure.

Very quick, described as speed "like an avalanche". Run-out


80 mm day prior, 160 mm in 5 days
Rhondda Main, na prior (between 1 to 5 year antecedent na
travelled down steep slopes, crossed railway line and
SW14 16 Feb 1928 na Failed four years after completion of tipping. Siddle et al (1996)
South Wales terminated at Ogwr Fawr river. Damaged railway lines,
event over 25 days).
telegraph poles and stone wall embankment.

Displacement of 3.5 m observed evening prior to


Tip located over spring associated Large noise. Movement described as "speed of Failure occured 1 year after abandonment of tip.
Pentre, South Lower portion of tip located on pre- failure, some 5 hours later failure occurred. Prior Siddle et al (1996),
SW15 4 Feb 1909 with outcrop of coal seam in None for 17 days prior to slide an avalanche". No confinefinement and minimal Suggestion of movement of pre-existing landslide caused
Wales existing landslide. to this, "sliding and giving way" observed during Knox (1927)
sandstone. obstruction. Deposition depth 3 to 4 m at toe. observed movement on evening prior to flowslide.
active tipping.
Suspect high foundation pore None for 2 days prior to slide, 80 mm on Day 3. Bulge or "boot" over 200 m length of toe area Movement described as "at terriffic speed". Instability reported as a problem throughout history of spoil
Cefn Glas, South pressures following period of heavy Day 3 to Day 17 heavy rainfall, approx 20 to 50 Tip located over pre-existing landslide.
SW16 3 Jan 1925 suggests slow movement over broad area prior Possibly obstructed by railway embankment heap. Failures observed within other areas of spoil heap. Siddle et al (1996)
Wales rain. year antecedent rainfall event. to flowslide over 50 m width. on lower slopes. Active tipping (?) at time of failure.

Possible movement of pre-existing landslide may have


Possible high foundation pore Rear scarp of 1965 failure coincident with larger Failure bowl almost fully evacuated suggesting
triggered rapid failure. Difficult to evaluate mechanism,
SW17 Parc, South Wales 1965 pressures as indicated by outburst na Tip located on pre-existing landslide. scarp associated with movement of pre-existing substantial outburst of water. Flowed down Siddle et al (1996)
potential out-burst, but this may have occurred subsequent
of groundwater. landslide. hillslope and terminated at Nant Cwn Parc River.
to the failure. Tip active at time of failure.
Tip covered boggy areas and stream courses. Flowed 50 m down-slope to river where deflected, then Failure confined to tip area that covered stream and
150 mm on day prior to failure.
Fernhill, South Large issue of water emerged from toe of spoil. flowed as debris flow or hyper-concentrated streamflow springs. Difficult to evaluate mechanism, potential out-
SW18 4 Dec 1960 Equivalent to approx. 50 yr antecedent na na Siddle et al (1996)
Wales Failure occurred in area where tip covered stream downstream. Eventually blocked culvert. No spoil remained burst, but this may have occurred subsequent to the failure.
event over 1 to 25 days prior to slide.
and springs. in failure bowl. Tip active at time of failure.

Brownie (FCL), na na na Near 10 days of 0.9 m/day displacement Normal mobility event. Small sturzstrom. na Golder Assoc (1995)
BC1 11/6/83
British Columbia
Brownie (FCL), Normal mobility event. Ran over Event BC1 Probable cause steep natural topography,
BC3 21/9/84 na na na na Golder Assoc (1995)
British Columbia debris. Distal part spread. aggravated by precipitation
Appendix B Page B16

Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles (Sheet 3 of 6)
Estimated Volume
Travel Classification Slope and Travel Geometry Width (m) Area (sq.m.)
(cu.m.)
Date of Initial Slide
No. Site / Stockpile Cross-section S/pile Travel
Failure Classification Height, H Length, L Length, Alpha 1 Alpha 2 Alpha 3 Alpha 4 Initial / Slide Deposit
Classification Geometry / Height, H s Distance H/L Initial Slide Flow Run-out
(m) (m) L toe (m) o (deg.) (deg.) (deg.) (deg.) Failure Area Area
Plan Shape (m) Angle ( )
Brownie (FCL), Flow slide, confined flow slide /
BC5 24/7/84 Confined, curved 470 260 1125 780 22.6 0.42 38 na 8 28 300000 - 100 250 34320 -
British Columbia translational debris flow

Brownie (FCL), Flow slide, Partly confined,


BC6 16/8/85 flow slide 360 215 950 610 20.7 0.38 38 na 11 24 140000 - na na -
British Columbia translational linear

Brownie (FCL), Flow slide,


BC7 29/6/85 flow slide Partly confined 360 200 580 310 31.6 0.62 38 na 32 27 50000 - ? na -
British Columbia translational (sliver)

2-Spoil (FCL), Flow slide,


BC9 1/30/84 flow slide 255 60 32.0 0.62 38 na - 10 775000 - 220 na -
British Columbia translational

Brownie (FCL), Flow slide,


BC11 19/6/85 flow slide Unconfined, linear 450 260 910 560 26.2 0.49 38 na 12 25 225000 - 100 34320 -
British Columbia translational

13 Seam (FCL), Flow slide,


BC12 14/8/86 flow slide Unconfined, linear 250 180 550 210 24.4 0.45 38 na 0 25-45 250000 - 100 150 23760 -
British Columbia translational

Brownie (FCL), Flow slide,


BC13 23/11/86 flow slide 290 500 na 38 na - 28 45000 - na -
British Columbia translational

B-stone (FCL), Flow slide, confined flow slide /


BC14 10/5/86 Confined, curved 525 300 1250 750 22.7 0.42 38 na 13 29 115000 - 150 70 59400 18000
British Columbia translational debris flow

Blackrill (FCL), Flow slide, confined flow slide /


BC16 17/6/85 Confined, deflected 475 140 1120 880 22.9 0.42 38 na 10 33 55000 - 80 90 14784 -
British Columbia translational debris flow

2-Spoil (FCL), Flow slide,


BC19 5/5/72 flow slide Unconfined, linear 275 120 650 460 22.9 0.42 38 na 8 25 50000 - 90 180 14256 -
British Columbia translational

2-Spoil (FCL), Flow slide,


BC20 27/5/72 flow slide Unconfined, linear 270 135 630 480 23.2 0.43 38 na 8 25 160000 - 300 280 53460 -
British columbia translational

2-Spoil (FCL), Flow slide,


BC24 8/11/74 flow slide Unconfined, linear 270 220 770 550 19.3 0.35 38 na 3 31 450000 - 180 220 52272 -
British columbia translational

2-Spoil (FCL), Slump (flow


BC27 12/11/74 flow slide Unconfined, linear 250 110 520 220 25.6 0.48 38 na 7 20 750000 - 200 180 29040 55000
British columbia slide ?)

H2 B3 (BLM), Flow slide,


BC35 18/11/80 flow slide Unconfined 260 230 450 160 29.9 0.58 38 na 5 35 na - 75 na -
British Columbia translational

A40-C1 (BLM), Flow slide, confined flow slide /


BC36 29/6/82 Confined, deflected 415 150 1370 1140 16.8 0.30 38 24 12 30 650000 - 200 200 39600 45600
British Columbia translational debris flow

5975 (BLM), Flow slide, Partly confined,


BC38 8/2/76 flow slide 335 260 880 550 20.8 0.38 38 na 7 26 na - 80 40 27456 25000
British Columbia translational deflected

#2 (BLM), British Flow slide,


BC40 24/11/68 flow slide Unconfined, curved 330 225 850 620 21.2 0.39 38 na 0 32 150000 - 100 na 90000
Columbia translational

A29-N (BLM), Flow slide, confined flow slide /


BC42 5/5/71 Confined, deflected 565 260 3000 2600 10.7 0.19 38 10.5 3 45 to 28 2200000 - 300 150 102960 450000
British Columbia translational debris flow

H Knob (BLM), Flow slide, confined flow slide /


BC43 13/5/71 Confined, deflected 870 120 3360 3200 14.5 0.26 38 16 (24 to 9) 9 36 na - 120 120 19008 -
British Columbia translational debris flow

A29-S (BLM), Flow slide, confined flow slide / 13.5 (25 to


BC44 25/5/73 Confined, deflected 500 120 2750 2500 10.3 0.18 38 3 42 110000 - 160 na -
British Columbia translational debris flow 8)
1640 MM (QCL), Flow slide, Partly confined,
BC48 21/6/86 flow slide 290 200 680 400 23.1 0.43 38 9.5 - 46 600000 - 100 140 26400 105000
British Columbia translational deflected

1705 WN (QCL), Flow slide, confined flow slide / 15.5 (20 to


BC49 30/6/86 Confined, deflected 500 210 1790 1750 15.6 0.28 38 6 (9 to 3) 46 to 20 1600000 - 150 200 41580 91000
British Columbia translational debris flow 9)

Flow slide,
1660 N (QCL), translational confined flow slide /
BC54 9/9/85 Confined, curved 515 240 2560 2200 11.4 0.20 38 9 (5 to 15) 4 20 (up to 40) 2500000 - 350 300 to 500 95040 735000
British Columbia (along fill / natural debris flow
interface)

1966 W (LINE), Flow slide, confined flow slide / 14.5 (30 to


BC62 1/7/82 Confined, deflected 490 136 2200 2000 12.6 0.22 38 9 to 5 31 250000 - 90 180 16157 80000
British Columbia translational debris flow 10.5)

East (GRH), British Flow slide, rotational Partly confined,


BC75 20/3/83 flow slide 360 160 1000 850 19.8 0.36 38 na run-up 28 400000 - 180 250 38016 -
Columbia / translational deflected

East (GRH), British Partly confined,


BC76 11/5/83 Flow slide, rotational flow slide 225 170 680 380 18.3 0.33 38 na 7 14 720000 - 250 175 56100 68400
Columbia linear

East (GRH), British Flow slide, Partly confined,


BC79 12/7/86 flow slide 200 160 780 530 14.4 0.26 38 na 5 18 480000 - 150 150 31680 24750
Columbia translational linear

East (GRH), British Flow slide, Partly confined,


BC80 12/2/87 flow slide 250 119 750 580 18.4 0.33 38 21.5 0 50 580000 - 200 200 31416 -
Columbia translational linear

East (GRH), British Flow slide, Partly confined,


BC81 7/3/87 flow slide 170 85 415 290 22.2 0.41 38 19 2.5 55 to 19 285000 - 150 150 16830 25000
Columbia translational deflected

North (GRH), Flow slide, confined flow slide /


BC89 1/7/85 Confined, deflected 360 200 1600 1200 12.7 0.23 38 na 6 24 300000 - 350 160 92400 45000
British Columbia translational debris flow
Appendix B Page B17

Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles (Sheet 4 of 6)

Date of
No. Site / Stockpile Hydro-geological Features Rainfall / Snowmelt Morphological Features Pre-Failure Deformation Behaviour Post-Failure Observations Comments References
Failure

Brownie (FCL), Groundwater seen flowing from the Normal mobility event. Toe debris = fine
BC5 24/7/84 na na na Failure did not reach crest. Golder Assoc (1995)
British Columbia toe a few days prior to failure. waste and till

Brownie (FCL), Normal mobility event. Initially gradual, toe


BC6 16/8/85 na na Very steep upper foundation na na Golder Assoc (1995)
British Columbia crept along creek.

Brownie (FCL), High wireline rates of movement and sliver Normal mobility event. Run-out data
BC7 29/6/85 na na na na Golder Assoc (1995)
British Columbia failures all month is estimated

2-Spoil (FCL), Pore water pressures under frozen


BC9 1/30/84 na na na Normal mobility event. na Golder Assoc (1995)
British Columbia crust

Brownie (FCL), Normal mobility event. Gradual slip over


BC11 19/6/85 na na na na na Golder Assoc (1995)
British Columbia top of existing debris

13 Seam (FCL), Inactive for 3 months. Toe not involved in


BC12 14/8/86 Groundwater seepage on face. na na na Normal mobility event. Failed over 30 mins. Golder Assoc (1995)
British Columbia failure.

Brownie (FCL), Crest unaffected. Inactive for 2 weeks.


BC13 23/11/86 na Snow melt. na na Normal mobility event. Very little data. Golder Assoc (1995)
British Columbia High load rate.

B-stone (FCL), Normal mobility event. 1990 survey shows Failure did not involve toe. Inactive for 4
BC14 10/5/86 na na na na Golder Assoc (1995)
British Columbia distal mound. months. Affected by blasting ?

Blackrill (FCL), Normal mobility event. Only 10000 cu.m.


BC16 17/6/85 na na na Suspect bulging face. Delayed failure. Rapid. Golder Assoc (1995)
British Columbia moved down channel.

2-Spoil (FCL),
BC19 5/5/72 na Very heavy snowfall na na Normal mobility event. Groundwater not felt important. Golder Assoc (1995)
British Columbia
2-Spoil (FCL), Normal mobility event. Over-rode haul road
BC20 27/5/72 na Very heavy snowfall na na Groundwater not felt important. Golder Assoc (1995)
British columbia embankment.

2-Spoil (FCL),
BC24 8/11/74 na na na na Normal mobility event. Rapid failure. Golder Assoc (1995)
British columbia
2-Spoil (FCL), Organic foundation soils. Organic Normal mobility event. Gradual,
BC27 12/11/74 na na na na Golder Assoc (1995)
British columbia soil displaced ahead. not dynamic run-out.

H2 B3 (BLM), Abnormal behaviour on 13 Sept, 2 months Normal mobility event. Probably a slump
BC35 18/11/80 na na na na Golder Assoc (1995)
British Columbia prior to failure. over old slide debris.

A40-C1 (BLM), Normal mobility event. Super-elevated.


BC36 29/6/82 na na na na Rapid failure, run-out well documented. Golder Assoc (1995)
British Columbia Good account of run-out.

5975 (BLM), na na na na Normal mobility event. Sharp turn. Limited data available. Golder Assoc (1995)
BC38 8/2/76
British Columbia
#2 (BLM), British na heavy rain Steep foundation slope. na Normal mobility event. Rapid. Triggered by heavy rain onto snowpack. Golder Assoc (1995)
BC40 24/11/68
Columbia
A29-N (BLM), High mobility event. Relatively small
BC42 5/5/71 Pore pressures in foundation. na Clay till foundation. No warning (?reported). na Golder Assoc (1995)
British Columbia volume ran out ?

H Knob (BLM), High mobility event. Exposed debris not Run-out was mudflow following earlier dump
BC43 13/5/71 na na na na Golder Assoc (1995)
British Columbia mud. Distal portion = fan. slump.

A29-S (BLM), High mobility event. Unusually large run-out


BC44 25/5/73 na na Steep toe topography. na na Golder Assoc (1995)
British Columbia distance. Due to organics?

1640 MM (QCL), Normal mobility event. Not rapid. Fine


BC48 21/6/86 Foundation pore pressures?. na na Toe bulge. High load rate over past 3 months. Golder Assoc (1995)
British Columbia carbonaceous waste in debris.

1705 WN (QCL), High mobility event. Flowed like water.


BC49 30/6/86 na na na na na Golder Assoc (1995)
British Columbia Explosive failure.

During dumping - 0.02 to 0.04 m/hour on average, High mobility, major event. High velocity as indicated by
Failed portion of tip located in gully Active at time of failure. Failed within dump due to wet
Would expect percolation through dump face to peak rates up to 0.1 m/hour. No monitoring of partial super-elevation. Failure contained in broad valley
area. Relatively steep slopes to fines. Suspect rainfall acted as trigger to failure and Golder Assoc (1995)
1660 N (QCL), concentrate in gully region. Possible stream at toe 86 mm in 40 days prior, 30 mm failure. Large cracks observed several hours down to RL 1225, thereafter valley narrowed. Coarser waste
BC54 9/9/85 sides of gully. Run-out within topography below dump concentrated flow to gully region. Dawson (1994)
British Columbia given relatively high rainfall. Suspect rainfall trigger in 6 days prior. prior to failure, dumping stopped at this point. carried on liquefied layer of finer sandy gravel. Liquefaction
confined gully. Organic soils Deposition on full length of run-out, indications of flow Dawson et al (1998)
to failure. Major cracking and bulging not evident until just of organic foundation below RL 1225 suspected to have led
within lower part of run-out. sliding on gradually depleting layer of finer sandy gravel.
before failure. to high mobility.

1966 W (LINE), High mobility event. Wet foundation major


BC62 1/7/82 na na Steep toe slope. na na Golder Assoc (1995)
British Columbia cause, not sturzstrom.

East (GRH), British Normal mobility event. Run-up on opposite


BC75 20/3/83 na na Steep toe slope. na Poor waste. Over-steep face. Golder Assoc (1995)
Columbia slope. Not sturzstrom.

East (GRH), British Excess foundation pore pressures. na na na Normal mobility event. Not sturzstrom. Not sturzstrom. Poor waste. Golder Assoc (1995)
BC76 11/5/83
Columbia
East (GRH), British Normal mobility event. Run-out over May 1985
BC79 12/7/86 Foundation seepage. na Steep topography. na Poor quality debris at toe. Golder Assoc (1995)
Columbia slide debris.

East (GRH), British Poor quality debris at toe. Affected


BC80 12/2/87 na na Steep terrain. na Normal mobility event. Golder Assoc (1995)
Columbia by previous slip?.

East (GRH), British Poor quality debris at toe. Frozen


BC81 7/3/87 na na na na Normal mobility event. Not a major rapid event. Golder Assoc (1995)
Columbia foundation pushed ahead.

North (GRH), Normal mobility event. Foundation organic soil


BC89 1/7/85 na na Steep foundation. na Sliver failure. Poor rock. High load rate. Golder Assoc (1995)
British Columbia pushed 350 m ahead.
Appendix B Page B18

Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles (Sheet 5 of 6)
Estimated Volume
Travel Classification Slope and Travel Geometry Width (m) Area (sq.m.)
(cu.m.)
Date of Initial Slide
No. Site / Stockpile Cross-section S/pile Travel
Failure Classification Height, H Length, L Length, Alpha 1 Alpha 2 Alpha 3 Alpha 4 Initial / Slide Deposit
Classification Geometry / Height, Hs Distance H/L Initial Slide Flow Run-out
(m) (m) Ltoe (m) o (deg.) (deg.) (deg.) (deg.) Failure Area Area
Plan Shape (m) Angle ( )
1690 MT (QCL), Flow slide, Unconfined,
BC112 17/8/84 flow slide 390 250 1070 740 20.0 0.36 38 na 9 17? 1500000 - 240 400 79200 135000
British Columbia translational deflected

1690 MT (QCL), Flow slide,


BC113 4/10/85 flow slide Unconfined, linear 440 250 1480 1180 16.5 0.30 38 na 6 21 2200000 - 250 300 82500 240000
British Columbia translational

Blaine (FCL), Flow slide,


BC149 16/6/86 flow slide Unconfined, linear 250 230 510 210 26.0 0.49 38 na 11 15 na - Full na -
British Columbia translational

SS STG 2 (FCL), Flow slide, confined flow slide /


BC153 29/5/90 Confined, linear 300 130 670 500 24.1 0.45 38 25 13 35 250000 - 200 50 34320 -
British Columbia translational debris flow

SS STG 2 (FCL), Flow slide, confined flow slide /


BC154 7/7/90 Confined, curved 350 140 870 690 21.9 0.40 38 21 (25 to 14) 14 49 300000 - 110 70 20328 -
British Columbia translational debris flow

Flow slide,
South (FCL), British translational (with Unconfined, linear
BC155 26/10/89 flow slide 490 410 1270 710 21.1 0.39 38 15 -24 (run-up) 33 to 24 3000000 - 230 500 to 600 127512 360000
Columbia rotational back- (flow spread)
scarp)

North (GRH), Flow slide, confined flow slide / Confined, deflected


BC156 22/11/89 translational debris flow (?)
400 260 2230 1950 10.2 0.18 38 4 - 48 to 4 1500000 - 350 na 190000
British Columbia
1660 (QCL), British Flow slide, confined flow slide /
BC157 7/11/87 Confined, deflected 490 280 2480 2200 11.2 0.20 38 8 5 47 to 8 5600000 - 640 200 236544 450000
Columbia translational debris flow

1680 WW (QCL), Flow slide, confined flow slide /


BC158 13/5/88 Confined, deflected 405 240 1110 790 20.0 0.36 38 14 9 36 80000 - ? 100 na 120000
British Columbia translational debris flow

Brownie (FCL), Flow slide, confined flow slide /


BC162 June 1983 Confined, curved 400 170 1050 870 20.8 0.38 38 (33 to 10 ?) (33 to 10 ?) 33 na - 60 na -
British Columbia translational debris flow

South (FCL), British Flow slide, 0 to -11.5


BC163 31/5/93 flow slide Unconfined, linear 375 385 1030 500 20.0 0.36 38 14 24.5 8000000 - 500 500 254100 600000
Columbia translational (run-up)

Flow slide
Cougar (GRH), translational confined flow slide /
BC164 11/5/92 Confined, curved 260 95 825 720 17.5 0.32 38 16 8.5 21 (up to 48) 160000 - 150 130 to 70 218381 55000
British Columbia (along fill / natural debris flow
interface)

2158 (LINE), British confined flow slide / 17.5 (27 to


BC165 24/3/90 Flow slide Confined, curved 308 160 920 660 18.5 0.33 38 9 to 5 31 100000 - 100 100 21120 -
Columbia debris flow 14)
2158 (LINE), British confined flow slide /
BC166 2/5/90 Flow slide Confined, curved 328 160 1340 1080 13.8 0.24 38 20 (27 to 14) 10 30 150000 - 100 100 21120 -
Columbia debris flow

Flow slide,
Peak Downs, Hay Unconfined, suspect
HP1 30 Aug 1974 translational on flow slide 11 11 45.7 30 13.5 0.24 35 (approx) 0 - 0 3500 - 55 - - -
Point spreading
basal interface

Flow slide,
Saraji (2 South), Unconfined, suspect
HP2 2 Aug 1974 translational on flow slide 9 9 26.9 14 18.5 0.34 35 (approx) 0 - 0 5000 - 120 - - -
Hay Point spreading
basal interface

Flow slide,
Saraji (1 South), Unconfined, suspect
HP3 9 Aug 1976 translational on flow slide 9 9 39.9 27 12.7 0.23 35 (approx) 0 - 0 5600 - 140 - - -
Hay Point spreading
basal interface

Flow slide,
translational on
Peak Downs (6 Unconfined, suspect
HP4 3 Feb 1977 basal interface flow slide 10.5 10.5 78.0 63 7.7 0.13 35 0 - 0 16000 - 150 170 - 12000
South), Hay Point spreading
(some retrogression
?)
Flow slide,
Goonyella (1 Unconfined, suspect
HP5 9 Mar 1977 translational on flow slide 10.5 10.5 45.0 30 13.1 0.23 35 (approx) 0 - 0 11300 - 200 - - -
North), Hay Point spreading
basal interface
Flow slide,
Saraji (1 South), Unconfined, suspect
HP6 9 Mar 1977 translational on flow slide 10.5 10.5 45.0 30 13.1 0.23 35 (approx) 0 - 0 10900 - 200 - - -
Hay Point spreading
basal interface
Flow slide,
Saraji (1 South), Unconfined, suspect
HP7 9 Oct 1978 translational on flow slide 13.5 13.5 43.3 24 17.3 0.31 35 (approx) 0 - 0 4500 - 60 - - -
Hay Point spreading
basal interface

Flow slide,
Peak Downs Unconfined, suspect
HP8 19 Dec 1974 translational on flow slide 6 6 23.2 13 14.5 0.26 35 (approx) 0 - 0 1250 - 22 - - -
(Expt), Hay Point spreading
basal interface

Flow slide,
Peak Downs Unconfined, suspect
HP9 19 Dec 1974 translational on flow slide 6 6 20.2 10 16.5 0.30 35 (approx) 0 - 0 850 - 15 - - -
(Expt), Hay Point spreading
basal interface
Appendix B Page B19

Table B1.5: Case studies of “rapid” landslides in coarse, granular stockpiles (Sheet 6 of 6)

Date of
No. Site / Stockpile Hydro-geological Features Rainfall / Snowmelt Morphological Features Pre-Failure Deformation Behaviour Post-Failure Observations Comments References
Failure

1690 MT (QCL), na na na na Normal mobility event. Not a major run-out event. Regressive with several distinct slumps. Golder Assoc (1995)
BC112 17/8/84
British Columbia
1690 MT (QCL), Normal mobility event. Toe pushed till slabs
BC113 4/10/85 na Heavy precipitation in last 2 months. na Previous slip. Poor rock. Golder Assoc (1995)
British Columbia ahead (previous event).

Blaine (FCL), Normal mobility event. Gradual. Stopped Frozen foundation ? Or strain induced pore
BC149 16/6/86 na na na na Golder Assoc (1995)
British Columbia against toe berm. pressures.

SS STG 2 (FCL), Pore pressures in colluvium? na Steep foundation. na Normal mobility event. Distal portion deflected. Fines in waste. Golder Assoc (1995)
BC153 29/5/90
British Columbia
SS STG 2 (FCL), Excess pore pressures in colluvium
BC154 7/7/90 Wet weather Steep toe. na Normal mobility event. Ran over previous debris. na Golder Assoc (1995)
British Columbia due to wet weather.
Normal mobility event. Much of flow slide debris crossed
Failure zone confined to section of From extensometers: Active dumping up until 24/10/89 when stopped due to
Intermittent stream in Blackwood Gully. Would creek and ran-up opposite slope. Large spread of material, Golder Assoc (1995)
South (FCL), British Significant in 50 days prior to slide spoil pile within Blackwood Gully. 24/10/89 - 0.1 m/hour excessive rates of movement. Finer sandy gravels
BC155 26/10/89 expect percolation through dump face to up to 600 m wide. Very rapid - air blast damage to nearby Dawson (1994)
Columbia (71 mm), none for 4 days prior. Steep, colluvial foundation slopes, 25/10/89 - accel. Rapidly associated with base of slide. Sharp contact between slide
concentrate in Blackwood Gully. foliage. Deposition - thin blanket on hill slope, most in base Dawson et al (1998)
up to 27 degrees. up to 1 to 1.3 m/hour prior to failure base and underlying foundation.
of valley.

North (GRH), Excess pore pressures in organic High mobility event. Unusually flat run-out.
BC156 22/11/89 na na na na Golder Assoc (1995)
British Columbia foundation. Thick organic soil.

1660 (QCL), British High mobility event. Run-out channeled in


BC157 7/11/87 na na na na na Golder Assoc (1995)
Columbia stream bed over previous failure debris.

1680 WW (QCL), na Normal mobility event. Rapid, fluid like. Crest steepened to 42 degrees. Dump fines. Golder Assoc (1995)
BC158 13/5/88 Pore pressure in colluvium. na na
British Columbia
Brownie (FCL), na Normal mobility event. na Golder Assoc (1995)
BC162 June 1983 na na Steep foundation.
British Columbia
South (FCL), British Pore water pressures in surficial Normal mobility event. Subsequent debris
BC163 31/5/93 na na na na Golder Assoc (1995)
Columbia foundation materials. slump down valley.

Very steep (up to 48 deg) on upper slopes


Buried snow and ice lenses observed Poor material - pre-dominantly fine materials,
of dump, approx 20.5 deg at toe of dump. Relatively narrow confined flow event. Golder Assoc (1995)
Cougar (GRH), in cuts of unfailed portion of dump. High snowmelt ? - several days of Cracks observed in dump face, defined by surficial soils (till and colluvium) and near
BC164 11/5/92 Dump located at head of broad gully. Suggestion of very rapid movement. Dawson (1994)
British Columbia Heavy rains during dumping. warm weather prior to slide. snowfall. No monitoring prior to failure. surface rock. Fatality. Dump inactive for 13
Foundation comprised Deposition along total length of run-out. Dawson et al (1998)
Snowmelt possibly triggered failure. months prior to failure.
thin layer of sand and gravel colluvium.

2158 (LINE), British na na Steep foundation. na Normal mobility event. sharp bend. Poor waste. 1 m/day crest advance. Golder Assoc (1995)
BC165 24/3/90
Columbia
2158 (LINE), British Normal mobility event. Ran over Event BC165
BC166 2/5/90 na na na na na Golder Assoc (1995)
Columbia debris.

Peak Downs, Hay Loose stockpile placed by stacker or dozer Placed very wet (dozer bogging in Layer 2),
HP1 30 Aug 1974 na Heavy at mine, 3.3 mm at Hay Point na na Eckersley (1986)
Point on prepared, relatively flat platform. moisture content = 11 to 16.7%

Loose, windrowed stockpile placed Windrows dozed over edge of pile. Placed
Saraji (2 South), Water level in stockpile 1.5 to 2 m 11 mm in period 6 to 9 days prior to Failure occurred some 2 days after completion
HP2 2 Aug 1974 by dozer on prepared, relatively flat na slightly wetter than usual. Failure occurred Eckersley (1986)
Hay Point height above base at failure. slide. of windrowing
platform. some 2 days after completion of windrowing.

Loose stockpile placed by stacker Stacking and dozing windrows at time of failure, windrows
Saraji (1 South), Water level in stockpile 0.3 to 1 m 3 Eckersley (1986)
HP3 9 Aug 1976 na and dozer on prepared, relatively na na dozed over edge of pile. Dry density ~ 0.79 to 0.92 t/m for
Hay Point height above base at failure Eckersley (1985)
flat platform. windrowed coal, 10 to 11.5% mc at Saraji

Water level in stockpile 1.2 to 1.8 m


Peak Downs (6 272 mm in 5 days prior to slide, most Loose stockpile placed by stacker Londitudinal cracks observed on face of Failure initiated near dumping point. Runout Stacking at time of failure. Loose tipped from Eckersley (1986)
HP4 3 Feb 1977 height above base at failure. Heavy
South), Hay Point of rainfall within first 30 hours. on prepared, relatively flat platform. stockpile prior to flowslide. obstructed by low height embankment. stacker, mc = 9 - 12 %. Eckersley (1985)
rainfall prior to failure.

Heavy rainfall prior to failure. Suspect Loose stockpile placed by stacker No stacking or dozing at time of failure. Occurred within 1 -
Goonyella (1 338 mm in 8 days prior to slide, Eckersley (1986)
HP5 9 Mar 1977 wet coal on placement and possibly and dozer on prepared, relatively na na 2 mins of HP6 slide. Failure most likely within windrowed
North), Hay Point 189 mm on day prior to slide. Eckersley (1985)
high water level in stockpile. flat platform. pile dozed over edge.
Heavy rainfall prior to failure. Suspect Loose stockpile placed by stacker No stacking or dozing at time of failure. Occurred within 1 -
Saraji (1 South), 338 mm in 8 days prior to slide, Eckersley (1986)
HP6 9 Mar 1977 wet coal on placement and possibly and dozer on prepared, relatively na na 2 mins of HP5 slide. Failure most likely within windrowed
Hay Point 189 mm on day prior to slide. Eckersley (1985)
high water level in stockpile. flat platform. pile dozed over edge.
Loose, windrowed stockpile placed
Saraji (1 South), No stacking or dozing at time of failure.
HP7 9 Oct 1978 na 6.1 mm, blown onto side of stockpile by dozer on prepared, relatively flat na na Eckersley (1986)
Hay Point Windrows dozed over edge of pile.
platform.

Loose stockpile dumped from low


Peak Downs drop height on prepared, relatively na na
HP8 19 Dec 1974 No stacking or dozing at time of failure. Placed
(Expt), Hay Point Articial injection of water to induce Artificial and natural of 113 mm, plus flat platform. with small drop, no compaction. Average dry
failure. Water level in stockpile 1.75 m inflow from wells, over 7 days prior Eckersley (1986)
density ~ 0.88 t/m3. Moisture at placement
max. height above base at failure to slide. Loose stockpile dumped from low
Peak Downs 8 to 13% (average 11%).
HP9 19 Dec 1974 drop height on prepared, relatively na na
(Expt), Hay Point flat platform.
Appendix B Page B20

Table B1.6: Flow slides from tailings impoundments – cases initiated by static liquefaction.
o
Failure Details Embankment Properties Ground Slopes ( ) Other Material Properties
Operating
Tailings Dam Failure Distance of Travel State at Under Down-
Height, Length, Length, Velocity Volume Travel Height Slope Construction Material Hydro-geological Features
Site Date Retrogression Distance 3 3 Time of o Type Dam stream Slimes Foundation
H (m) L (m) L toe (m) o (m/s) (10 m ) Classification (m) () Method Properties
(m) Angle ( ) Failure (Alpha 4) (Alpha 2)

2.7 to 11 slurry flow initially, SM - silty fine sand


4 km to river, 1.3 ML - clayey silt (99% Clay (CH) highly
Bafokeng, South 47 (est to 4900 (to (10 on developed into hyper- Ring Dyke, Upstream, (13% fines) with very Lot of water ponded on dam prior to
11/11/74 45 km to 1000 0.55 3000 in operation 20 33 thin slimes layers (kh
1.3 reducing to fines, 18% CF), strongly reactive, highly
Africa river) river) exiting concentrated stream Platinum mine Mechanical rainfall.
storage dam 0.29 contractant fissured (Black Clay)
dam) flow >> kv )

Foundation groundwater levels


highest in June - July following
CL/CI interlayered with snowmelt and heavy rain. Water
200 8.5 quickly developed into in operation Centreline then SM - silty fine to
SM. 85-100% fines (CF
Recent slope deposits
4400 to Sidehill, 12 (10 to level in tailings dam reasonably high
Stava, Italy 19/7/85 400 (est) 4400 (virtually nothing 5.20 increasing 190 hyper-concentrated (on and off for 29.5 34 upstream, medium sand (10 to 4.5 to 7.6 20 - 33%), IL ~1.2 (0.9 to
(soft) over glacial
River Avisio Fluorite mine 18) as water continually on dam.
left) to 12 to 25 stream flow 23 years) cyclone 60% fines) deposits
1.9). Change in pond operation resulted in
high phreatic surface in sand
embankment.

Seepage constantly observed at


ML - fine sandy silt down-stream toe, drains not
slurry flow initially, Layering of tailings, Aeolian silty sand to
(60 to 80% fines, CF < functioning. Large quantity of water
Merriespruit, poss. developed into in operation Ring dyke, finer layers within 1.5 m deep over
22/2/94 74 2330 2000 330 1.83 na 600 31 24 Upstream (?) 0 1.5 10%), soft to stiff stored in impoundment (no return
South Africa hyper-concentrated (for 16 years) gold mine coarser more sandy clayey sand/silty clay
(Su = 34 kPa avg), water dam). Pool located adjacent
stream flow materials (kh >> kv) alluvium (soft ?).
PI = 1 to 8%. to embankment that failed, limited
freeboard.
Clays and Sandy
Ring Dyke, High groundwater level in tailings.
Upstream, Sandier portion of Clays - stiff, over-
in operation Phosphate ML - fine sandy silt Seepage exited at toe of slope rather
Texas, USA 1966 10 410 300 110 1.40 2.5 to 5 105 slurry flow 11 26.5 Mechanical, tailings from upper 0 0 consolidated
(for 4 years) mine (gypsum (60% fines), non plastic. than collected in drains (sand drains
under-drainage. beach deposition. Pleistocene marine
tailings) did not function).
clays

Arcturus, Ring dyke,


1978 33 390 300 90 4.84 na 20 slurry flow in operation 25 46 Upstream (?) na 1.5 1.5 na na na
Zimbabwe gold mine

Saaiplass -
Ring dyke, -0.5
Slide 1, South 3/18/92 13.5 255 80 175 3.03 na 60 slurry flow in operation 21.2 21 Upstream (?) na -0.5 na na No water on dam prior to failure.
gold mine (upslope)
Africa
Saaiplass -
Ring dyke, -0.5
Slide 2, South 19/03/92 12.2 160 70 90 4.36 na 120 slurry flow in operation 18.6 21 Upstream (?) na -0.5 na na No water on dam prior to failure.
gold mine (upslope)
Africa

Saaiplass - Ring dyke,


Slide 3, South 22/03/92 31.2 525 300 225 3.40 na 150 slurry flow in operation 26 23 Upstream (?) na 1 1 na na na
gold mine
Africa

SM - gravelly silty 0.56 to Dam No. 3 stored very large volume


Dam across Dumped coarse CL/ML - sandy silt (54%
- hyper-concentrated in operation 4.8 (up sand (13 to 20% Man (0.82 Slimes/Sludge under of water, almost overtopping prior to
Buffalo Creek 26/2/72 322 27600 27.2 km 0.67 2.8 to 4.6 420 58 valley, coal coal waste over 2.5 fines), discharge from
(not flow slide) stream flow (for 2 years) to 38) fines), coarse coal reducing to Dam Nos 2 and 3. failure. High water table through
mine existing sludge plant
waste. 0.42) embankment.
Appendix B Page B21

Table B1.6: Flow slides from tailings impoundments – cases initiated by static liquefaction (Sheet 2 of 2)
Tailings Dam Failure Failure Mechanics (Trigger and Post-Failure Characteristics
Rainfall Pre-Failure Observations Post-Failure Deformation Behaviour Comments References
Site Date Mechanism) (Channelisation/Obstructions/Deposition)

Spread to width of 800 m a short distance beyond the Channelised flow within broad, flat gully and river.
8:45 am - overtopping and breach of diagonal Minimal ground erosion. Flow eventually stopped Fourie (2000)
Piping caused initial embankment failure. embankment, then entered broad flat channel and flowed
wall. 10:15 am - jet of water from face of once entered water storage 45 km downstream from Destructive flow slide, some 13 people died. Blight (1997)
Bafokeng, South 75 mm on night Once breached, tailings exposed and (at quick rate) some 4 km to river. (At river slide reported
11/11/74 embankment. Cracks developed and blocks fell 6 Horseshoe shaped post-failure shape in Jennings (1979)
Africa before failure liquefied. Suspect significant amount of to be 800 m wide and 10 m deep). Flow continued down tailings dam. 1 x 10 deposited between dam and
out of wall above water jet. Breach widened impoundment with bottle neck through breach. Rudd (1979)
erosion due to water volume.
rapidly to 100 m.
river (suspect turbulent, extremely quick flow) for some 40 river, 2 x 10 6 flowed to water storage. 10 m depth Midgley (1979)
km to a water storage impoundment. tailings near river.

Initial movement possibly as laminar flow. Suspect Channelised flow for most of length. Broad flow to
High phreatic surface within steep sand shell Jan 1985 ( 6 months prior) - small slip in Upper Massive catastrophe, some 268 people killed. Blight (1997)
1985 - wetter than developed reasonably quickly into a turbulent stream flow Stava (200 to 250 m wide) and then narrow width of
of Upper Basin caused initial drained failure. Basin and development of sinkhole. Lagoons Greater velocity of slide on flatter angle below Chandler and Tosatti (1995)
Stava, Italy 19/7/85 usual, 170 mm in 7 once entered stream some 50 to 100 m below dam. flow in river channel (50 to 100 m) to River Avisio.
Exposed tailings liquefied and flowed over- back in operation 4 days before failure. Lower Stava suggests increasing fluidity of tailings flow. Morgenstern (2000)
weeks prior to failure. Thren developed into very rapid and destructive Terminated at confluence of Stava tributary and
topping Lower Basin. pond virtually full prior to failure. Very little of tailings dam remained. Berti et al (1988)
hyperconcentrated stream flow. River Avisio. No information on deposition.

Over-topping, following high rainfall and Toe sloughing and excessive seepage during life
Castasrophic event. Tailings flowed through
exaserbated by limited freeboard. Erosion of dam, toe buttress unsuccessful in preventing Fourie (2000)
Merriespruit, 50 mm in 0.5 hour Flow stopped when entered ornamental lake 2 km village of Merriespriut causing destruction, some
22/2/94 initiated instability resulting in retrogressive sloughing. Slips on lower slope of dam several Suspect as quite a rapid flow. Blight (1997)
South Africa preceeding failure. from initial breach. 17 people died. Series of stepped terraces within
failure of the embankment. Exposed tailings days prior to failure. Series of bangs heard at Wagener et al (1998).
impoundment, bulb shaped.
liquefied and flowed. time of initial embankment failure.

Seepage at toe caused softening and minor


drained failures (also piping). These Flow analysis by Jeyapalan et al (1983b). Kleiner (1976)
Average velocity estimated from 1 to 2 minute event time. No channelisation or obstruction. Depostion along full
Texas, USA 1966 na developed into a larger drained failure of na Reasonable results obtained. Series of stepped Jeyapalan et al (1983a)
Flow spread broadly on flat slopes beyond breach. runout length (2 m thick some 50 m from the breach).
embankment. Exposed tailings liquefied and terraces within impoundment, bulb shaped. Jeyapalan et al (1983b).
flowed.

Initial drained failure of dam, probably due to


Arcturus, Heavy rain Flow stopped by low embankment of tailings return
1978 raised phreatic surface following heavy rain. na na 1 person killed. Blight (1997)
Zimbabwe preceeded failure. water reservoir.
Exposed tailings liquefied and flowed.

Saaiplass - Not clear, possibly just drained failure of Within impoundment - series of stepped terraces
Slide 1, South 3/18/92 no rain prior to slide downstream slope. Exposed tailings na na Flow obstructed by embankment. Blight (1997)
separated by steep low height sections.
Africa liquefied and flowed.
Saaiplass - Not clear, possibly drained failure of
Slide 2, South 19/03/92 no rain prior to slide downstream slope of embankment. Exposed na na Flow obstructed by embankment. na Blight (1997)
Africa tailings liquefied and flowed.
Not clear, possibly drained failure of
Saaiplass - embankment due to high phreatic water table
19 mm rain prior to Within impoundment - series of stepped terraces
Slide 3, South 22/03/92 related to rapid rise of tailings level in month na na Flow obstructed by embankment of return reservoir. Blight (1997)
slide separated by steep low height sections.
Africa preceeding failure. Exposed tailings liquefied
and flowed.
Dam 3 in meta-stable state. High phreatic 2 hrs prior - cracks across front of Dam 3, crest
Not a flow slide. Comprised a series of rotational
surface caused initial failure. Subsequent had sunk in some places. Just prior to slide - Turbulent flow at 2.75 to 4.6 m/sec down river. Water Confined flow within narrow valley and narrow gully.
89 mm in 2 days failures followed by breach and large release of W.A. Wahler & Associates
retrogressive failures due to high preatic front of had dam broken off, settlement and volume = 530,000 m3, soilds = 170,000 m3 beyond Dam Deposition along entire valley to town of Man (27 km
Buffalo Creek 26/2/72 leading up to failure, water (530,000 cu.m.) that eroded slimes. (1973)
surface and seepage from high up on face. d/stream displacement of embankment 1. Most material did not travel far. Of material below downstream), most of deposition within first 5 km
1 in 2 to 3 year event. Classify as turbulent or hyperconcentrated Johnson (1980).
Last part failed violently releasing stored observed. Then large part of central section Dam 1 - 75000 erosion of refuse bank and 23000 sludge. beyond Dam 1.
water. broke loose. streamflow.
Appendix B Page B22

Table B1.7: Flow slides from tailings impoundments – cases initiated by dynamic liquefaction.
Earthquake Details Failure Details Embankment Foundation
o
Date of Slope ( ) Properties of Slimes Failure Scarp Within
Tailings Dam Dist. to Operating State Pre-Failure Observations Post-Failure Observations Comments References
Failure Length, Volume Height Slope Construction (Alpha 4 & (Source Area) Impoundment
Magnitude Epicentre 3 o Material Properties at Time of Comments
L toe (km) (1000 m ) (m) ( ) Method Alpha 2)
(km) Failure
Turbulent flow down Barahona Creek
Rolled earth starter dam to and Cachapoal Rvr for 50 km. Very
Upstream, Silty Sand (SM) - fine to Near vertical frontal scarp at Catastrophic event, 54 persons
Barahona 1, operating (for 11 7m height then upstream 8.5 deg. immed. Failed 2 to 3 minutes after fast, wave climbed as high as 60 m Troncoso (2000),
1/10/28 8.2 95 50 3000 65 27 gravitational with medium sand, 15 to 26% na breach, series of stepped killed. Crest of dam 17 m
Chile years) construction. 15,500 down-slope. earthquake. on bends. Sand volcanoes on terraces Troncoso et al (1993)
separator cones fines terraces back into impoundment. above slimes.
tonnes/day deposit rate. and springs daylighting in scarps
post-failure.

Near vertical frontal scarp.


Silty Sand (SM) - fine to emergency storage Clayey Silt (ML) - low to medium
El Cobre- Old, Upstream, Since Dec 1963 only used for Dust raised by earthquake. Several Horseshoe shaped shell left. Catastrophic event, killed Troncoso (2000),
28/03/65 7.5 40 12 1900 35 34 medium sand, 45% fines , (operating 4 plasticity, 93% fines, wL = 19 to na
Chile gravitational emergency storage chunks of dry crust left on terraces. Central depression formed by more than 200 people. Dobry & Alvarez (1967)
non-plastic 35 years) 48%, PI = 0 to 19%, IL = 1.2
several almost horizontal terraces.

Silty Sand (SM) - fine to Very rapid deposition of Clayey Silt (ML) - low plasticity, Earthquake liquefied slimes. Waves Tailings flowed through the breach and
El Cobre- New, Upstream, operating (for 1.25 Troncoso (2000),
28/03/65 7.5 40 12 500 19 15 medium sand, 30% fines , tailings, approx. 2000 4 90% fines, wL = 27%, PI = 4 to produced on surface of pond. ran 8 m up the mountain in front, then na Tailings dam totally destoyed.
Chile cyclones years) Dobry & Alvarez (1967)
non-plastic tonnes/day. 5%, I L = 1.0 to 1.5 Erosion or slip in corner of dam. flowed down the valley.

Silty Clay (CH) - high plasticity, Dams located on gentle fluvial


Hierro Viejo, Upstream, Sandy Silt (ML) - fine Series of 5 dams, 1 in use 100% fines, wL = 55%, PI = Crust broke into blocks that floated on a Terraced slope within depression terraces of Petorca River (within Troncoso (2000),
28/03/65 7.5 18 1 1 5 45 operating 1 to 2 na
Chile gravitational sand, 62% fines, non-plastic while others dry out. mass of liquefied tailings. left after flow. valley floor). Possibility of Dobry & Alvarez (1967)
31%, I L = 0.83
some confinement of tailings flow.

Dam located in narrow valley with


Dam 3 in use at time of Clayey Silt (ML) - low plasticity,
Los Maquis gravitational Silty Sand (SM) - fine operating (for 28 Slimes transformed into liquid mass Suspect flowed within confined, steep 15 to 20 degree slopes. Personnel Troncoso (2000),
28/03/65 7.5 15 5 25 15 30.5 earthquake. Production 15 to 20 100% fines, wL = 35%, PI = 9%, na
No. 3, Chile (upstream) sand, 46% fines, non-plastic years) before the earthquake ended. valley. heard sound like an explosion and Dobry & Alvarez (1967)
approx. 35 to 45 tonnes/day. I L = 1.11
then the sound like a waterfall.

Production approx. 170 Silty Clay to Clayey Silt (CL/ML) Semi-circular front slope of 7 to 10
La Patagua - gravitational Sandy Silt (ML) - fine Troncoso (2000),
28/03/65 7.5 22 5 40 15 35 operating tonnes/day. Lot of water on 3 to 5 - medium plasticity, 100% fines, na na m height. Behind scarp several na
New, Chile (upstream) sand, 55% fines, non-plastic Dobry & Alvarez (1967)
dam at time of failure. wL = 43%, PI = 18%, IL = 0.99 step-like terraces left.

Slimes transformed into liquid mass.


Low and semi-circular frontal
Silty Clay to Clayey Silt (CL/ML) Waves formed on the surface of
Cerro Negro Upstream, Sandy Silt (ML) - fine operating (for Production approx. 240 scarp. Within impoundment Troncoso (2000),
28/03/65 7.5 38 5 95 20 40 20 - medium plasticity, 100% fines, the pond which eroded the small na na
No. 3, Chile gravitational sand, 52% fines, non-plastic 4 years) tonnes/day. terraced levels with intermediate Dobry & Alvarez (1967)
wL = 47%, Pi = 18%, I L = 0.99 dyke and tailings poured out of
small scarps.
the breach.

Scarps behind crest developing 3 - Vertical horseshoe shaped vertical


Clayey Silt (ML) - low plasticity, Dry upper crust cracked and floated in
gravitational Silty Sand (SM) - fine Production approx. 70 to 75 4 months prior to failure, increasing scarp of 7 to 15 m height. Troncoso (2000),
Bellavista, Chile 28/03/65 7.5 55 0.8 75 20 30.5 operating 7 87% fines, wL = 26%, Pi = 3 to chunks on liquefied mass. Sands left as na
(upstream) sand, 43% fines, non-plastic tonnes/day of tailings. in rate leading up to failure. Toe Terraced levels within central Dobry & Alvarez (1967)
4%, IL = 0.98 lateral ridges.
bulging. depression.

Clayey Silt (ML/MH) - medium Tailings dam built on on


Ramayama No. 1, gravitational Silty Sand (SM) - fine Material flowed down to the foot of the Troncoso (2000),
28/03/65 7.5 85 na 0.15 5 33 operating na 30 to high plasticity, 100% fines, wL na na mountainside with 30 degree
Chile (upstream) sand, 25% fines, non-plastic mountain. Dobry & Alvarez (1967)
= 49%, PI = 18%, IL = 1.14 slope.

Gold mine. 16 m high starter


~20 ( steep Sandy Clayey Silt (ML) - low Flowed down mountainside into a stream
No information. 3 m high dam then upstream
Mochikoshi Dam operating (for 38 mountainside), plasticity, 85% fines (CF < and eventually into the ocean. 1 km Travel angle approx. 1.2 degrees
14/1/78 7.0 40 (approx.) 30 80 32 15.5 upstream dyke constructed each construction (after 1965). na na Marcuson et al (1979)
No. 1, Japan years) then < 0.5 deg 10%), wL = 20 to 30%, PI = 2 to downstream flow depth approx. (H ~ 630 m, L ~ 30 km).
year. 120 tonnes/day deposition
rate. to ocean (?) 13%, IL = 1.4 1 to 1.5 m depth.

Gold mine. 8 m high starter


~20 ( steep Sandy Clayey Silt (ML) - low
dam then upstream
Mochikoshi Dam 5.7 close to dam No information. 3 m high operating (for 38 mountainside), plasticity, 85% fines (CF < Failure some 5 hours after
15/1/78 3 22 17.5 upstream construction (after 1965). na na na Marcuson et al (1979)
No. 2, Japan (aftershock) site (?) dyke constructed each year. years) then < 0.5 deg 10%), wL = 20 to 30%, PI = 2 to aftershock.
120 tonnes/day deposition
to ocean (?) 13%, IL = 1.5
rate.

Veta del Agua 2,


7/11/81 6.5 85 na na 20 40 Upstream na operating na na na na na na na Troncoso (2000)
Chile

Earthquake caused slimes to


Breach allowed liquefied tailings to flow.
Veta del Agua 1, Upstream / liquefy. Embankment failes with No under-drainage within Troncoso (2000),
3/3/85 7.8 80 8 na 24 26.5 to 34 na operating na 10 na Emptied approx. 40 % of the imponded na
Chile Centre line an "explosive burst" in the embankment. Troncoso (1988)
tailings.
central, highest region of the dam.

After failure - deep gorge within crest of


Cerro Negro 4, Upstream / Silty sands (SM), non- No under-drainage within Troncoso (2000),
3/3/85 7.8 105 10 100 40 22 operating Cyclone na na na dyke, approx 30 m wide, through which na
Chile Centre line plastic embankment. Troncoso (1988)
tailings and water flowed.
Almendro, Chile 14/10/97 7.0 100 na na 18 34 Upstream operating na na na na na na na Troncoso (2000)

Algarrobo, Chile 14/10/97 7.0 80 na na 20 34 Upstream operating na na na na na na na Troncoso (2000)


operating /
Maiten, Chile 14/10/97 7.0 120 na na 15 34 Upstream na na na na na na na Troncoso (2000)
abandoned
La Cocinera, Upstream /
14/10/97 7.0 80 na na 30 30.5 abandoned na na na na na na na Troncoso (2000)
Chile Centre line
Appendix B Page B23

Table B1.8: Case studies of flow slides in sensitive clays


Slope, Failure and Travel Geometry
Slide Slide Material Description Pre-Failure Post-Failure
Date of Initial Slide Slope Distance of Failure Travel Basal Volume Hydro-geological Velocity
Site Height, Length, Length, Alpha 1 Alpha 2 3 Area Width (Liquefied Material on Rainfall Failure Mechanics Deformation Deformation Comments References
Failure Classification Height, Retrogression, L (m) Surface Distance H/L plane (m ) Features (m/s)
H (m) L (m) toe (deg.) (deg.) (ha) (m) Surface of Rupture) Behaviour Behaviour
H s (m) Re (m) Depth (m) Angle (o ) (deg.)
Flow slide, Sensitive clay, soft to firm Initial failure triggered retrogressive Edgars and Karlsrud (1983)
Hekseberg na 40 na 700 160 540 12 3.27 0.057 na na na 1.8E+05 na na S t = 50 to 150 na na na na na
Retrogressive (?) sliding Trak and Lacasse (1996)

Instability of steep river bank due to Blocked Furre Namsen Translational slide, limited
Flow slide, rising (and high) groundwater. Active channel. Obstructed by retrogression. Quick clay overlain
Silty Clay - low plasticity, firm, High groundwater levels
translational erosion an aggravating factor (ice jam island and passive by glacial silts and sands of low
Furre, Norway 14/4/59 18 23 350 300 10 to 150 24 2.94 0.051 24 0 to -5 4 to 6 2.8E+06 18 850 highly sensitive (quick) (higher than usual in na na 0.5 to 1.3 Hutchinson (1961)
(some in February resulted in rapid erosion at pressures associated with sensitivity. Suspect translational
PI = 10%, St > 70, I L = 2.1 normal winter)
retrogression) toe of initial slide). Occurred as a slide surface location below component due somewhat to soil
translational "flake slide". base of channel. profile.

Sensitive clay, low plasticity,


Flow slide, soft to firm, PI = 3 to 5%, Initial failure triggered retrogressive Edgars and Karlsrud (1983)
Verdal na 70 na 9100 2000 7100 30 0.44 0.008 na na na 5.5E+07 na na na na na na 10 to 15
Retrogressive (?) sliding Trak and Lacasse (1996)
St = 20 to 200, IL = 2.5
Sensitive clay, low plasticity,
Flow slide, Initial failure triggered retrogressive Edgars and Karlsrud (1983)
Selnes na 25 na 450 140 310 10 3.18 0.056 na na na 1.4E+05 na na soft, PI = 3 to 9%, na na na na na
Retrogressive (?) S t > 100, I L = 1.9 sliding Trak and Lacasse (1996)

Above normal
Initial instability of river bank due to
Lemieuse (Sth snowfall in winter,
rising groundwater levels from rainfall Small slumps
Flow slide, Clay - high plasticity, firm, Suspect high heavy falls in Flowed down South Nation
Nation Rvr), 20/6/93 na na na 680 na 8 to 23 na na na na na 2.8E+06 17 na and snowmelt. Followed by series of observed in week na Lawrence et al (1996)
Retrogressive PI = 40%, St = 10 to 100 groundwater level March and April. River.
Ottawa retrogressive failures triggered by prior to main failure
Heavy Spring
continued back-scarp instability.
rain.

Initial instability of river bank due to


Suspect high
rising groundwater levels from rainfall
South Nation Flow slide, Clay - high plasticity, firm, groundwater level Flowed down South Nation Lawrence et al (1996)
16/5/71 na 24.5 na 490 na 24.5 na na 10 na na 6.0E+06 28 750 Wet Spring and snowmelt. Followed by series of na na
River Retrogressive PI = 40%, St = 10 to 100, I L = 1 following heavy River. Eden et al (1971)
retrogressive failures triggered by
snowmelt.
continued back-scarp instability.

Clayey Silt - low plasticity, Deep translational slide. Slide


firm, high sensitivity (quick) Instability due to rising groundwater Edgars and Karlsrud (1983)
Flow slide, Suspect high within quick clay below approx. 20
Baastad, Norway 5/12/74 34 24 1200 175 1000 24 1.62 0.028 21 1 2 1.5E+06 8 480 PI = 4 to 8%, S t = 50 to 100, Wet Autumn following rainfall. Earthworks was an na Channelised na Trak and Lacasse (1996)
translational groundwater level m of low sensitivity (IL < 1) silty
aggravating factor. Translational slide. Gregerson and Loken (1979)
IL = 1.6 clay.

Initial instability due to earthworks,


Flow slide, which triggered relatively slow,
Sensitive clay, low plasticity, Slow retrogressive slide followed Edgars and Karlsrud (1983)
translational, retrogressive sliding due to continued Rapid movement of
Rissa, Norway 29/4/78 35 25 1100 1350 ? 10 1.82 0.032 6 na 5 5.5E+06 33 800 very soft, PI = 5 to 12%, na na na 2.8 to 5.5 by quick translational slide. Initial Trak and Lacasse (1996)
some back-scarp instability. Followed by translational slides.
S t ~ 100, I L = 2.3 slide Ns = 7.4 (R/D = 75). Gregersen (1981).
retrogression several very large, retrogressive,
translational "flake" slides

Flow slide, Sensitive clay, low plasticity, Initial failure triggered retrogressive Edgars and Karlsrud (1983)
Skjelstadmarken na 45 na 2800 600 2200 15 0.92 0.016 na na na 2.0E+06 na na na na na na 0.2
Retrogressive (?) firm, PI = 6 to 10%, IL = 1.3 sliding Trak and Lacasse (1996)

Sensitive clay, low plasticity,


Flow slide, soft, PI = 6 to 8%, Initial failure triggered retrogressive Edgars and Karlsrud (1983)
Borgen na 29 na 1500 165 1350 9 1.11 0.019 na na na 1.2E+05 na na na na na na na
Retrogressive (?) sliding Trak and Lacasse (1996)
St = 20 to 100, IL = 1.4

Initial instability of over-steepened stream


Silty Clay - low plasticity, High groundwater bank due to rising groundwater. Active Rate of retrogression = 3.75 m/sec
cemented, highly sensitive Edgars and Karlsrud (1983)
following thaw and erosion of stream bank an aggravating Small failure in bank Channelised flow down (early stages 0.15 m/sec).
St. Jean Vianney, Flow slide, Tavenas et al (1971)
4/5/71 73 24 3880 550 3330 22.5 1.08 0.019 45 na na 6.9E+06 26.8 370 (quick), soft to firm (Champlain snowmelt resulting in na factor. Followed by series of of Petit Bras 10 days stream and river for 7.2 Occurred in crater of previous
Quebec Retrogressive Clay) PI = 7 to 12%, Trak and Lacasse (1996)
partial flooding. First retrogressive failures due to continued before main slide. some 3 km. slide. Samples indicate clay to be
Bentley and Smalley (1984)
S t > 200, I L = 1.8 heavy rain after thaw. back-scarp instability that progressed dilatent.
quickly in latter stages of slide.

Initial instability in steep river bank due


Possible high
to rising groundwater. Active erosion
Clayey Silt - stiff, low groundwater table Slide more than four
La Grande River, Flow slide, suspected as an aggravating factor. Flowed into La Grande
5/5/87 na 70 na 290 na 42 na na 35 na 6 3.5E+06 12 550 plasticity, sensitive, following snowmelt and na months prior to this na Lefebvre et al (1991)
Quebec Retrogressive Followed by series of retrogressive River.
PI = 8%, IL = 1.8 thaw. Strong failure.
failures due to continued back-scarp
underdrainage in till.
instability.

Translational slide, limited


Instability of river bank due to rising (and
Clayey Silt - low plasticity, High groundwater levels. Occurred rapidly. Dammed retrogression. Quick clay overlain
high) groundwater levels following wet
Flow slide, Mild and very wet Wet February river. Obstructed by by glacial silts and sands of low
Vibstad, Norway 22/2/59 16.3 28 510 260 250 25 1.83 0.032 22 0 to -3 2 to 9 1.0E+06 7 370 firm, sensitive, PI = 8%, and mild late Winter. Active erosion an na na Hutchinson (1965)
translational IL = 1.3 February. Ground not (mild) opposite bank of river and sensitivity. Suspect translational
aggravating factor. Occurred as a
frozen and saturated also river itself. component due somewhat to soil
translational "flake slide".
profile.

Slide triggered by embankment Single compound slide, no


Flow slide, Silty Clay / Clayey Silt - low construction works located at the rear retrogression. Triggered by fill
plasticity, very soft, highly Moved 15 to 20 m in 15 to
Bekkelaget, translational, Hydrostatic groundwater of the actual slide. Occurred as single placement at rear slope. Suspect
7/10/53 16 12 160 140 20 6 5.71 0.100 2 2 1 1.0E+05 1.6 155 na na 20 seconds as single 1.0 Eide and Bjerrum (1956)
Norway rotational back- sensitive (quick), PI = 9%, from ground surface failure with frontal section translational translational front due somewhat
St = 80, IL = 2.4 failure.
scarp at shallow depth and rotational rear to passive failure following failure
component. at rear.

Initial instability in road embankment with


Silty Clay - high plasticity, soft, Intense rainy period in stream at toe possibly triggered by high
Flow slide, sensitive, S t = 15 to 40, Autumn (Sept - Nov), groundwater levels and/or active stream Rate of retrogression avg 2.5
Tuve, Sweden 30/11/77 17 na 850 320 530 10 to 15 1.15 0.020 na 0.54 na 4.0E+05 28 190 na na na na Duncan et al (1980)
Retrogressive probable artesian erosion at toe. Followed by series of m/sec
IL = 1.3
pressures rapid retrogressive failures due to
continued back-scarp instability.
Appendix B Page B24

Table B1.9: “Rapid” landslides in submarine slopes (Edgars and Karlsrud 1983)
Landslide Geometry
Failure Travel Velocity Volume Material
Site Height, Length, Length, 3 Trigger Comments References
Date Distance H/L (m/s) (m ) Description
H (m) L (m) Ltoe (km) o
Angle ( )
Initial failure occurred at time of
exceptionally low tide.
Glacial - loose Mechanism - collapse and flow Karlsrud & Edgars (1982)
Orkdalsfjord, Possibly small slide in man-
2/5/30 500 22500 > 20 km 1.3 0.022 7.0 2.5E+07 sand, silt and clay of loose sand and silt Edgars and Karlsrud (1983)
Norway made fill
deposits. sediments. Travel distance Terzaghi (1956)
and rate determined by broken
submarine cables.

Earthquake or rapid
Bassien Ancient 2200 215000 200 + km 0.58 0.010 na 8E+11 na Edgars and Karlsrud (1983)
sedimentation

Storegga Ancient 1700 160000 160 + km 0.61 0.011 na 8E+11 na Earthquake (?) Edgars and Karlsrud (1983)

Grand Banks 1929 5000 750000 750 + km 0.38 0.007 7.7 7.6E+11 Sand / silt Earthquake Edgars and Karlsrud (1983)

Gravelly clayey
Spanish Sahara Ancient 3100 700000 700 km 0.25 0.004 na 6E+11 rapid sedimentation Edgars and Karlsrud (1983)
sand

Walvis Bay (SW


Ancient 2100 250000 250 km 0.48 0.008 na 3E+11 na na Edgars and Karlsrud (1983)
Africa)

Icy Bay /
Ancient 80 12000 12 km 0.38 0.007 na 3.2E+10 Clayey silt Earthquake Edgars and Karlsrud (1983)
Malaspina

Copper River Ancient 85 8000 8 km 0.61 0.011 na 2.4E+10 Silt / sand rapid sedimentation Edgars and Karlsrud (1983)

Clayey and sandy rapid sedimentation


Ranger Ancient 800 37000 37 km 1.24 0.022 na 2E+10 Edgars and Karlsrud (1983)
silt (earthquake)
earthquake / rapid
Mid Alb. Bank Ancient 600 5300 5.3 km 6.46 0.113 na 1.9E+10 silty clay Edgars and Karlsrud (1983)
sedimentation

Wil. Canyon Ancient 2800 60000 60 km 2.67 0.047 na 1.1E+10 silty clay and silt rapid sedimentation Edgars and Karlsrud (1983)

Kidnappers Ancient 200 11000 11 km 1.04 0.018 na 8.0E+09 sandy silt, clay Earthquake (?) Edgars and Karlsrud (1983)

rapid sedimentation /
Kayak Trough Ancient 150 18000 18 km 0.48 0.008 na 5.9E+09 clayey silt Edgars and Karlsrud (1983)
earthquake

Paoanui Ancient 200 7000 7 km 1.64 0.029 na 1.0E+09 silt / sand Earthquake (?) Edgars and Karlsrud (1983)

Mid. Atl. Cont.


Ancient 300 3500 3.5 km 4.90 0.086 na 4.0E+08 silty clay rapid sedimentation Edgars and Karlsrud (1983)
Slope

Magdalena R. 1935 1400 24000 24 km 3.34 0.058 na 3.0E+08 na rapid sedimentation Edgars and Karlsrud (1983)

clayey and sandy


California Ancient 150 3500 3.5 km 2.45 0.043 na 2.5E+08 Earthquake (?) Edgars and Karlsrud (1983)
silt

Suva, Fiji 1953 1800 110000 110 km 0.94 0.016 na 1.5E+08 sand Earthquake Edgars and Karlsrud (1983)

Valdez 1964 168 1280 1.28 km 7.48 0.131 na 7.5E+07 gravelly silty sand Earthquake Edgars and Karlsrud (1983)

quick clay (?) and


Sokkelvik, Norway 1959 100 5000 5 km 1.15 0.020 na 5.0E+06 na Edgars and Karlsrud (1983)
sand

Sandnessjoen,
1967 180 1200 1.2 km 8.53 0.150 1 to 3 5.0E+05 na Blasting, Man made fill Edgars and Karlsrud (1983)
Norway

Helsinki Harbour,
1936 11 400 .4 km 1.58 0.028 na 6000 sand / silt Man made fill Edgars and Karlsrud (1983)
Norway
Appendix B Page B25

Table B1.10: Failure case studies of hydraulic fill embankment dams, sub-aqueous fills and Wachusett Dam.
Slope and Failure Geometry Embankment / Fill Description
Failure Initial Slide Travel Dam/Fill Travel Volume Width Foundation
Site Height, Length, Length, Alpha 1 Alpha 2 Alpha 4 Embankment / Fill
Date Classification Classification Height, Distance H/L (m3) (m) State (Age) Type Construction Description
H (m) L (m) Ltoe (m) (deg.) (deg.) (deg.) Material (Source Area)
H s (m) Angle (o )

Silty Sand / Sandy Silt


Flow slide, Completed 1917 Silty Sand / Sandy Silt
Sheffield Dam, Hydraulic Fill (SM/ML) - DI = 35 to
29/6/25 translational basal Flow slide 7.6 7.6 64 30 6.8 0.12 22 0 0 16000 90 (failed 8 years Hydraulic (SM/ML) - DR = 35 to 50% (75 -
USA Dam 40% (upper loose
component after construction) 80% SMDD)
layer)

Silty Sand / Sandy Silt


Flow slide, (SM/ML) - Clean sand toward
Lower San translational basal Completed 1925 outer part of shoulder, siltier
Hydraulic Fill Hydraulic, rolled Stiff clay with lenses of
Fernando Dam, 9/2/71 component, Flow slide 31 35 177 76 9.9 0.18 21.8 -2 to -3 -2 to -3 400000 400 (failed 46 years (SM/ML) to centre. Highly
Dam upper section. sand and gravel.
USA rotational back- after construction) stratified.
scarp. 12 - 85% fines. DR = 50 to
55%.

Flow slide,
Variable - outer section
translational basal
Calaveras Dam, Failed during Hydraulic Fill Hydraulic, sluiced gravelly grading to Silty Gravelly alluvium over
24/3/18 component, Flow slide 56 56 440 190 7.3 0.13 18 0 0 600000 215
USA construction Dam soft sandstone Sand/Sandy Silt (SM/ML) to bedrock
rotational back-
silt/clay in liquid core.
scarp.

Upstream
Flow slide, rotational Zoned shoulder Sand - Loose to very loose Dense to very dense
Wachusett Dam, Failed during first
11/4/07 (?), possibly Flow slide (?) 24.4 24.4 179 100 7.9 0.14 22 0 0 47000 213 Earthfill comprised fine sand with some gravel glacial silts, sands and
USA filling
retrogressive. Embankment dumped fill (no and silt, some silty sand. gravels.
compaction)

Alluvium - sands,
Flow slide, Sand (SP) - fine to medium
Fort Peck Dam, 520 to Failed during Hydraulic Fill gravels and clays of 0
22/9/38 translational basal Flow slide 58 58 880 510 3.8 0.07 12.6 0 0 8000000 Hydraulic sand grained, trace silt/clay. (e =
USA 800 construction Dam to 27 m depth over
component. 0.67 to 0.75). Stratified.
Bearpaw Shale.

July to
Nerlerk Berm - Flow slide / debris Failed during Hydraulic
August 27 27 625 375 2.7 0.04 11 0 to 0.2 0 to 0.2 80000 100
Slide 1 flow (?) construction Sea Berm
1983
Flow slide. Suspect Hydraulic, placed
July to Sand (SP) - fine to medium Thin layer of soft to
Nerlerk Berm - translational basal Flow slide / debris Failed during Hydraulic using special
August 27 27 550 400 2.6 0.05 11 0 to 0.2 0 to 0.2 100000 110 grained, some silt (2 to 15%). very soft clay (Su = 4 to
Slide 2 component and flow (?) construction Sea Berm nozzle to steepen
1983 D R = 25 to 35%. 7 kPa)
retrogressive. slopes, very loose
July to
Nerlerk Berm - Flow slide / debris Failed during Hydraulic
August 27 27 625 400 2.8 0.04 11 0 to 0.2 0 to 0.2 150000 150
Slide 4 flow (?) construction Sea Berm
1983
5 (18
Flow slide. Sand (SP) - medium to coarse
between
Translational basal Failed during sand, trace silt. Very loose Alluvial - thin layer
Willamette River, Flow slide / debris 28 (25 to hydraulic Hydraulic and
18/1/67 component between 20.7 20.7 72 38 14.0 0.29 1 4000 23 to 46 construction by River berm bulldozed (DI < 15%), organic clayey silt over
Terminal 2 flow 34) and bulldozed
dozed and hydraulic excavation at toe hydraulic loose to med dense sand.
bulldozed
fill, retrogressive. (DR - 20 to 60%)
fill)
Appendix B Page B26

Table B1.10: Failure case studies of hydraulic fill embankment dams, sub-aqueous fills and Wachusett Dam (Sheet 2 of 2)
Failure Pre-Failure Deformation Velocity
Site Hydro-geological Features Trigger Mechanism Post Failure Deformation Behaviour Comments References
Date Behaviour (m/s)

Flow slide on liquefied, loose silty sand / sandy silt base.


Liquefaction of saturated, loose to medium Earthquake triggered failure. Failure
Low phreatic surface downstream of Upper partly saturated and outer denser zones carried on
Sheffield Dam, Earthquake, Mag. 6.3, dense silty sands to sandy silts close to Seed et al (1969),
29/6/25 upstream facing section. na liquefied basal zone, only minor break-up of outer zone. na in downstream slope. Estimated
USA 11 km from dam. foundation / embankment interface Seed (1987)
Groundwater above foundation. Movement hinged about abutment. No channelisation or S ur = 2.5 kPa.
following earthquake.
obstruction. Deposition along full length of travel.

Liquefaction of saturated medium dense Flow slide on liqufied layer of SM/ML in base of Castro et al (1985),
Reservoir level 11 m below crest at Earthquake triggered failure. Failure
Lower San silty sands and sandy silts in lower section embankment. Initial large slide followed by several Seed et al (1975),
time of failure. Initial rise in 0.6 to 1 in upstream slope. Estimated Sur of Castro et al (1992),
Fernando Dam, 9/2/71 Earthquake, Mag. 6.6. of hydraulic fill following earthquake. na retrogressive slides. Outer zones remained intact (carried
piezometers observed in downstream Bazier and Dobry (1995)
USA Stratification may have had a significant on liquefied layer). No channelisation or obstruction. 19 to 24 kPa from dynamic analysis.
shoulder after earthquake. Gu et al (1993)
influence. Deposition along full length of travel.

"Abnormal" movement of upstream section


Low water level in impoundment. of embankment observed since June 1917 Material with greater than 45 to 50%
Suspect shear strain induced liquefaction Whole mass moved forward then broke up as front section
Suspect high pore pressures in liquid (9 months before failure). Construction Seed (1987),
Calaveras Dam, Construction. Long lead of weaker lenses of fill near base of dam. moved at greater rate. Steam shovelled fill remained intact voids flowed, 40% stable. Failure in
24/3/18 core, with phreatic surface in induced movement. Approx 1.2 m total 1 Hazen (1918),
USA up to failure. Suspect surface of rupture close to base (carried on liquefied layer). No channelisation. Deposition upstream slope. Estimated Sur of 36
upstream shoulder close to water horiz movement before failure. Day of Hazen & Metcalf (1918).
of embankment. along full length of travel, 6 to 25 m deep. kPa.
level in reservoir. failure - initially slow movement then
increased in rate.

Suspect near horizontal phreatic Failure in upstream slope on first


Suspect collapse of loose dumped sands Flow slide. Little information. Possibly retrogressive. No
Wachusett Dam, surface in upstream shoulder Rising water level in filling. Estimated Sur of 16 kPa
11/4/07 on wetting during filling triggered na channelisation. Deposition along full length of travel, 0 to 12 9 (?) Olson et al (2000)
USA commensurate with reservoir reservoir. based on kinetics (3.8 kPa lower
liquefaction. m deep. bound). Pre-failure, 38 to 42 kPa.
impoundment level.

Construction. Excess
Reservoir level dropped 5.8 m in 2 Toe bulging of upstream shoulder and Upper and outer part of embankment remained virtually Flow slide carried intact surface Middlebrooks (1940),
pore pressures in
months prior to slide. High core pool Suspect differential movements triggered "abnormal" subsidence movement of core intact (carried on liquefied layer). Sand boils still issueing 10
Fort Peck Dam, bentonitic seams in blocks. Suspect liquefaction of Casagrande (1965),
22/9/38 level. Suspect relatively flat phreatic liquefaction. Initiated at the upstream toe pool (detected hours prior to failure). days after slide. All over in 10 minutes. No channelisation, 2 to 5
USA foundation. Initial hydraulic fill, mainly in more silty ENR (1939a),
surface in upstream shoulder, of the core pool. Increase in rate of movement. Almost inlet channel may have obstructed travel. Deposition 0 to 5 stratified layers. Seed (1987).
movement on these
commensurate with reservoir level. suddenly rapid movements started. m at toe, 15 to 20 m near dam.
bentonite seams.
July to Failure in outer loose hydraulic fill
Nerlerk Berm - Collapse and flow down to slopes gradients of 1 to 2 na
August Saturated (sub-aqueous) na placed using new nozzle to steepen
Slide 1 Suspect collapse and static liquefaction of degrees. No channelisation or obstruction.
1983 slopes. Suspect failure triggered by
July to loose to very loose sand triggered by differential shear along interface
Nerlerk Berm - Construction, trigger Collapse and flow down to slopes gradients of 1 to 2 Sladen et al (1985b),
August Saturated (sub-aqueous) shear movement due to active fill na na between loose and medium dense
Slide 2 possibly localised over- degrees. No channelisation or obstruction.
1983 placement. Shear movements possibly sand. Possibly semi-continuous Sladen & Hewitt (1989),
steepening of loose to Rogers et al (1990)
along interface between loose to medium layers of loose sand rather than
very loose sand.
July to dense fill and foundation. Suspect mass loose sand. Estimated S ur of
Nerlerk Berm - retrogressive. Collapse and flow down to slopes gradients of 1 to 2 na
August Saturated (sub-aqueous) na 0.25 to 1 kPa using Stark and
Slide 4 degrees. No channelisation or obstruction.
1983 Mesri (1992).

Suspect triggered by shear strains along


Excavation at toe interface between bulldozed and hydraulic Occurred as a series of retrogressive rotational slips on a
Willamette River, resulting in over- fill as a result of effective stress changes translational interface between the bulldozed very loose fill Estimated S ur of < 1 kPa using Stark
18/1/67 Near saturated sand (sub-aqueous) na 0.3 to 0.5 Cornforth et al (1974)
Terminal 2 steepening (approx 34 associated with excavation at toe. and hydraulically placed fill. No channelisation or and Mesri (1992).
degrees) Retrogressive failure occurred along this obstruction.
interface.
Appendix B Page B27

2.0 MECHANICS OF SHEARING OF CONTRACTIVE


GRANULAR SOILS

2.1 INTRODUCTION

This section presents a more detailed review of the literature on several aspects of the
mechanics and behaviour of contractive granular soils on shearing that are only
summarily covered in Chapter 3 (Section 3.3) of the thesis. These aspects include:
• Empirical field methods for evaluation of liquefaction and flow liquefaction
potential from case studies of earthquake induced liquefaction;
• Assessment of the residual (undrained) shear strength of flow liquefied soils from
laboratory testing;
• The effect of fines content on liquefaction; and
• Discussion on the merits and pitfalls of laboratory testing.

2.2 EMPIRICAL FIELD METHODS FOR EVALUATION OF L IQUEFACTION AND FLOW


LIQUEFACTION POTENTIAL

The purpose of this section is to discuss in more detail the field methods used for
evaluation of cyclic liquefaction potential. The methods for evaluation of flow
liquefaction potential for statically triggered events are discussed in the thesis (Section
3.3) and not here.
Methods for evaluation of liquefaction potential have been developed from the
results of laboratory testing and empirically from field case studies. Most of these
methods are related to cyclically induced liquefaction (i.e. from earthquakes).
The early work of Harry Bolton Seed pioneered much of the development of
empirical methods for evaluation of liquefaction potential from case study analysis of
earthquake-triggered liquefaction. Empirical methods for evaluation of large
deformation potential have been developed for; the Standard Penetration Test (SPT), the
Cone Penetration Test (CPT), shear wave velocity from seismic refraction, and the
Becker Penetration Test (BPT), applicable to gravelly soils where the SPT is not
appropriate. Empirical correlations to SPT and CPT are the most widely used given the
larger field database of information. There are only a few sites where BPTs were
carried out and where liquefaction occurred (Youd et al 1997; Harder 1997), therefore it
Appendix B Page B28

cannot be directly correlated with liquefaction resistance and is typically used to


estimate an equivalent SPT.
Andrus and Stokoe (1997) present methods for evaluation of liquefaction potential
from shear wave velocity results. Youd et al (1997) comment that the limitations of
shear wave velocity methods are that the seismic wave velocity measurements are made
at small strains compared to the large strain phenomena of liquefaction, and that it
requires geotechnical investigation to recover physical samples for classification to aid
in liquefaction assessment. However, the latter limitation is also applicable to the CPT,
and the use of shear wave velocity in conjunction with SPT would presumably provide a
good means of investigation. Also, soils tend to fall into the “it may liquefy zone”
rather than the “it will” or “it will not” liquefy, and therefore, at present, shear wave
velocity does not provide a good method of separation.

2.2.1 SPT for Evaluation of Liquefaction Potential

Seed (1979) developed a correlation between the cyclic stress ratio to cause liquefaction
and the N value of the SPT test based on field data of sand deposits subject to strong
shaking from earthquakes. The case studies analysed were essentially level ground
conditions. The correlation proposed by Seed et al (1985), with mi nor adjustments
proposed by Youd et al (1997), is presented in Figure B2.1. This correlation of
liquefaction potential based on cyclic resistance is for large deformation and not
necessarily for flow liquefaction.
Seed et al (1983) suggested that a correlation of field performance to the cyclic
strain induced by earthquake might improve the basis for liquefaction assessment.
Figure B2.2 presents the Seed et al (1985) correlation, with slight modifications from
the 1996 NCEER workshop (National Centre for Earthquake Engineering Research), of
cyclic resistance ratio for clean sands incorporating consideration of the level of strain
associated with liquefaction (Robertson and Wride (1997) provide methods for
corrections of the SPT value for overburden pressure, energy ratio and other factors).
The line on Figure B2.2 for shear strains of 20% can be considered as representing a
boundary of flow liquefaction conditions.
With respect to the effect of fines content (non plastic fines), Seed et al (1985) found
that for a given corrected SPT value ((N1)60) the cyclic resistance of the soil increased
with increasing fines content. Robertson and Wride (1997) comment that it is not clear
Appendix B Page B29

whether the cyclic resistance of the soil is increased because of greater liquefaction
resistance or in fact the ((N1)60) value of the soil decreases as a consequence of the
increase in compressibility and decrease in permeability associated with increasing fines
content.

Figure B2.1: Field correlation for cyclic resistance ratio from SPT data (Seed et al
(1985) with minor adjustments by Youd et al (1997))

Fear and McRoberts (1995) reviewed the Berkeley data set upon which Seed based his
liquefaction assessment. They found that each case record was reduced to a single
corrected SPT value ((N1)60) reflecting the state of the cohesionless soil, and this was
not the minimum (N1)60 value. The significance of this from a design perspective was
that a conservative assessment of liquefaction potential would be obtained if the loosest
material on site were evaluated using the Seed criterion. Fear and McRoberts (1995)
evaluated the data set (including an additional 25 case records) by selecting the lowest
(N1)60 for each case record and developed a lower bound (based on non-liquefied cases)
and upper bound (based on liquefied cases) criterion (Figure B2.3). The classification
Appendix B Page B30

of cyclic liquefaction, the same as that by Seed for the Berkeley data, was based on
observation of surface features including sand boils, cracking, lateral spreading or
settlement. They found that the upper bound criterion, i.e. case studies of liquefaction
and marginal liquefaction, for clean sands was asymptotic to (N1)60 values in the range
13 to 15, much lower than those of the Seed criterion. The bounds for sand with fines
(Fear and McRoberts 1995) are at lower corrected SPT values than for clean sands,
indicating that these soils (sands with fines) are more resistant to liquefaction.

Figure B2.2: Cyclic resistance ratio (CRR) for clean sands under level ground
conditions incorporating consideration of potential strain associated with liquefaction
(Robertson and Wride 1997)

Fear and McRoberts (1995) commented that the major factors in evaluating the
liquefaction potential were the cyclic stress ratio of the earthquake, the corrected SPT
Appendix B Page B31

value and fines content. Other factors that influence the potential for liquefaction were
the topography and material properties including gravel content, cementation, aging,
fabric, layer thickness and drainage characteristics. They commented that the
impedance of upward drainage, such as by an upper soil cap of lower permeability, may
increase the soils susceptibility to liquefaction.

(a) (b)

Figure B2.3: Field correlations of cyclic stress ratio versus (N1)60 of (a) lower bound
from non-liquefied case records and (b) upper bound from liquefied and pressure relief
case records (Fear & McRoberts 1995).

2.2.2 CPT for Evaluation of Liquefaction Potential

Plots of cyclic liquefaction potential similar to those based on SPT results have been
developed for tip resistance from CPT results with similar type correction factors also
applicable. The reader is referred to Robertson and Wride (1997) and Olsen (1997) for
recent papers discussing correlations of CPT results for assessment of cyclic
liquefaction potential.
Appendix B Page B32

2.3 PROBLEMS WITH L ABORATORY TEST METHODS FOR EVALUATION OF FLOW


LIQUEFACTION POTENTIAL

Some of the issues relating to difficulties with laboratory test methods and interpretation
are discussed including; the effects of sample disturbance, laboratory assessment of
liquefaction, and problems with the use of laboratory prepared samples.

2.3.1 Sample Disturbance

There is significant debate over the merits of laboratory testing as a method of


evaluating the flow liquefaction potential of soils in the field. Small changes in void
ratio or soil structure due to sample disturbance can significantly affect the laboratory
results, and laboratory prepared specimens are not suitable for assessment of field
performance due to the effects of stratification and fabric on the results. At the National
Science Foundation workshop on the shear strength of liquefied soils (NSF 1998) the
working group agreed that high quality “undisturbed” samples are required for
laboratory testing if the results are to be used for assessment of field liquefaction
evaluation, and the general consensus was for ground freezing as the preferred method
of sampling. They recognised that ground freezing may not be suitable for silty sands
and sandy silts where the freezing action could cause sample disturbance.

2.3.2 Laboratory Assessment of Liquefaction

Ishihara (1993), for undrained cyclic laboratory tests, considered liquefaction to have
occurred when 5% double amplitude strain had been reached in cyclic stress triaxial
compressive tests and 3 to 4% in simple shear tests. Robertson and Wride (1997)
considered that this definition is conservative and in conflict with field observations,
citing the results of Yoshimi et al (1994) as evidence (Figure B2.4). As shown, the
results for frozen samples using the 5% double amplitude strain criterion indicate cyclic
liquefaction occurs for clean sands with (N1)60 values of up to 40, which is much higher
than that from field observations (Figure B2.1 and Figure B2.3) at similar (N1)60 values.
Robertson and Wride (1997) comment that for loose sands the strain criterion is
applicable as it “often occurs close to the point at which the effective confining is
essentially zero and deformations develop rapidly”. However, they consider that for
denser sands the strain criterion may be unduly conservative as “the 5% double-
Appendix B Page B33

amplitude strain can occur before sufficient pore pressure has developed to take the
sample to the state of essentially zero effective stress” and “deformations may actually
be progressing rather slowly”. Poulos (1998) discussed the need to define the term
“liquefaction” given their was no distinguishing between its use in field cases of
undrained stability failure and its definition as excessive strain in laboratory testing.

Figure B2.4: Comparison between liquefaction resistance and SPT (N1)60 values from
tube samples and undisturbed in-situ frozen samples (Yoshimi et al 1994)

2.3.3 Problems with the Use of Laboratory Prepared Samples For Field Assessment

Ishihara (1993) commented that samples prepared by different methods (moist tamped,
water pluviated or dry deposition) show different resistances to flow liquefaction due to
the difference in soil structure produced (Figure B2.5). Many authors (including Been
et al 1991; Poulos 1998) have discussed the significant effect of soil fabric on
liquefaction resistance. In general, they consider laboratory testing using prepared
samples only of use for evaluate of field conditions if the sample preparation methods
can replicate reasonably the in-situ soil structure, otherwise testing on high quality
undisturbed samples is required.
Appendix B Page B34

Figure B2.5: Effect of sample preparation method on isotropic consolidation (Ishihara


1993).

The work by Yoshimi et al (1994) compared laboratory test results from both ground
freezing and piston undisturbed sampling techniques (Figure B2.4) from a deposit of
clean sand. The results from in-situ samples taken after freezing reasonably replicate
the shape of the cyclic resistance curve developed from field case analysis. On the other
hand, the results from piston type undisturbed samples bear little resemblance. Ishihara
(1993) commented that laboratory results indicate piston sampling is useful for recovery
of undisturbed samples from loose sand deposits, but sample disturbance effects become
increasing pronounced as sand density increases.
Robertson and Wride (1997) and Ishihara (1993) considered the effects of aging and
fabric to have a significant effect on the soils liquefaction resistance and that laboratory
results from reconstituted samples would not reflect the true behaviour of the in-situ
sand. Both authors refer to the cyclic test results after Yoshimi et al (1989) comparing
samples from reconstituted freshly deposited sand to in-situ frozen samples (Figure
B2.6). The results indicate that the cyclic resistance of reconstituted fresh samples is
significantly less than that of the in-situ deposit. Whilst the effect of aging would be
difficult to quantify, intuitively it would be expected to have an effect on the results,
particularly in silty sands and sandy silts where cementation may occur.
Appendix B Page B35

Figure B2.6: Comparison of liquefaction resistance between freshly deposited and in-
situ undisturbed samples of sand (Yoshimi et al 1989)

2.4 CHARACTERISTICS OF CONTRACTIVE AND DILATIVE SOILS SHEARED UNDER


MONOTONIC LOAD CONDITIONS

2.4.1 Characteristics of Clean Sands

The purpose of this section is to summarise the characteristics of clean sand observed in
monotonic laboratory testing and discuss issues such as the uniqueness or otherwise of
the steady state line, effects of sample fabric and the effects of stress path based on the
published results of research in these areas. Most of the discussion is centred on
monotonic stress or strain controlled tests, which would describe more closely the
behaviour sands under static liquefaction conditions. The effects of fines content are
discussed in Section 2.4.2 and gravelly soils in Section 2.4.3.
Ishihara (1993) presented the results of laboratory undrained triaxial compression
tests on Toyoura sand, a fine to medium grained, sub-angular, uniform clean sand, to
describe the characteristics of clean sand. The samples were prepared by different
preparation techniques (moist placement, dry deposition and water sedimentation),
isotropically consolidated, and tested under strain controlled monotonic loading
conditions. The results indicate that the steady state condition in undrained loading
(Figure B2.7) is dependent on the void ratio alone.
For sands at an initial state in e log p' space (e = void ratio, p' = mean normal
effective stress) to the right of the steady state line, on shearing in drained conditions the
Appendix B Page B36

sand will contract toward the steady state line. In undrained loading conditions the
tendency for contraction is suppressed and post peak strain weakening is observed
(Figure B2.8) as load is transferred to the pore fluid and the effective stress path moves
to the left (at constant void ratio) toward steady state. For sands at an initial state on the
left side of the steady state line, when sheared in drained loading conditions will dilate.
In undrained loading conditions the dilation is suppressed and strain hardening is
observed (Figure B2.9) as pore water pressures reduce and load is transferred onto the
soil structure.

Figure B2.7: Steady state line of Toyoura sand (Ishihara 1993)

Figure B2.8: Undrained behaviour of states on the contractive side of the steady state
line (Ishihara 1993)
Appendix B Page B37

Figure B2.9: Undrained behaviour from states close to and on the dilative side of the
steady state line (a) stress-strain plot and (b) effective stress paths (Ishihara 1993)

A “quasi steady state” condition in undrained loading (where strain hardening follows
post-peak strain weakening) is observed for some initial states as shown in Figure B2.9.
It is generally observed where the initial state is slightly above or slightly below the
steady state. The change from strain weakening to strain hardening is termed the phase
transformation. Ishihara (1993) also proposed the concept of the initial dividing line,
IDL (Figure B2.10), below which strain hardening is observed and above which stain
weakening followed by strain hardening (or quasi behaviour) is observed. The
published test results of Figure B2.8 appear to contradict this definition of the IDL
because at the states defined by Points c, d, and e (Figure B2.10), all above the IDL, do
not show the quasi behaviour. Possibly the IDL is ill defined.
Appendix B Page B38

In several very loose samples, prepared by moist preparation, zero residual strength
was observed post peak (Figure B2.8). Ishihara (1993) defined the void ratio at or
above which samples exhibit zero residual strength as the threshold void ratio, eo
(Figure B2.10).
Ishihara (1993) showed the isotropic compression line, quasi steady state line and
initial dividing line are all dependent on the fabric or structure of the sample, void ratio
and initial effective confining stress conditions. The quasi steady state condition is not
therefore a fundamental state parameter for sands.

Figure B2.10: Initial states of moist-placed samples from Figure B2.8 and concept of
the threshold void ratio (Ishihara 1993)

When clean sands are tested in undrained conditions from the same initial void ratio and
different initial effective stress conditions the steady state strength reached for all
samples will be the same in accordance with the concept of uniqueness of the steady
state line (Figure B2.11).
Sladen et al (1985a) proposed the concept of the collapse surface (or instability line)
for evaluation of the liquefaction potential of very loose sands located on the contractive
side of the steady state line. They defined the collapse surface as the locus of the peak
deviatoric stresses of undrained triaxial compression tests of samples tested at different
initial mean normal stress from the same initial void ratio (Figure B2.11). On reaching
the collapse surface (in undrained conditions) the soil structure collapses and load is
transferred onto the pore fluid. The post collapse effective stress path is downward and
Appendix B Page B39

leftward toward the steady state line. The residual undrained strength, defined by the
steady state deviatoric stress, qss, is dependent only on the initial void ratio. An
important point here is that the strength defined by the instability line is lower than the
critical state strength, which is commonly used in stability analysis.

Figure B2.11: Collapse surface concept, locus of peak deviatoric stress of undrained
triaxial tests for samples prepared at the same initial voids ratio (Sladen et al 1985a).

Been and Jefferies (1985) defined the concept of the “state parameter”, f , to uniquely
characterise the state of a sand. Its validity is based on the concept that the steady state
line is a unique fundamental parameter, regardless of effective stress path. The state
parameter (Figure B2.12) is defined as the difference between the initial void ratio, e?,
and the void ratio at steady state corresponding to the initial mean normal effective
stress, ess (f = e? - ess). The concept of the state parameter is that it characterises the
effective stress path and normalised steady state strength of samples with identical
values of the state parameter, regardless of the initial mean normal effective stress. It
was initially developed for assessment of the characteristics of hydraulic fills, and Been
and Jefferies (1985) cautioned its usage for assessment of geologically aged sand
deposits.
Appendix B Page B40

Figure B2.12: Definition of the state parameter (Been and Jefferies 1985).

Yamamuro and Lade (1998) defined three “normal” states of behaviour for clean sands
(Figure B2.13) for samples (prepared at the same initial void ratio) tested undrained
under monotonic load conditions:
• Stable behaviour, where the undrained stress path is strain hardening. This is
similar to the behaviour observed for clean sands sheared from initial states well to
the left of the steady state line that are strongly dilative and representative of sands
with large negative state parameter, ϕ .
• Temporary instability, where phase transformation or a quasi steady state condition
occurs post the initial peak undrained strength. Decreasing stability is observed
with increasing confining pressure in this region.
• Instability (liquefaction), indicative of the contractive tendency resulting in the
development of positive pore pressures to a steady state condition. Once the
effective stress path exceeds the peak deviatoric stress the sand, in undrained
conditions, becomes unstable and failure occurs. This corresponds to states to the
right of the steady state line (Figure B2.7). Yamamuro and Lade (1998) comment
that this behaviour is observed at relatively high confining pressures, the high
confining pressure suppressing the tendency for dilation.

The above summary describes the characteristics of sand based on the concept that
the steady state line is a unique parameter for clean sand. However, the concept of
uniqueness of the steady state line is queried by the results of research, in particular
Appendix B Page B41

from research using stress paths other than those for drained and undrained compressive
testing and also from testing of sands with fines (Section 2.4.2). The results of research
on this aspect of uniqueness (or otherwise) are discussed below.

Figure B2.13: Schematic diagram showing different types of behaviour for undrained
stress paths of samples at the same void ratio and different initial confining pressures
(Yamamuro and Lade 1998).

Yamamuro and Lade (1997) considered that the “normal” behaviour of sand (described
above) did not adequately describe their results of laboratory undrained triaxial tests on
very loose sands and silty sands (Lade and Yamamuro 1997). They considered that five
different states existed for very loose sands and silty sands (Figure B2.14). In addition
to the states described earlier, Yamamuro and Lade (1998) consider that for very loose
sands (and silty sands) the additional two states are (Figure B2.14):
• Liquefaction region, which occurs at low confining stresses and is characterized by
large pore water pressure development resulting in zero effective stresses at low
axial strain levels (stress path AO). Increasing confining pressure results in
increasing effective stress friction angle at peak undrained strength in this stress
region.
• Temporary liquefaction region, characterized by stress path BCDE, which occurs at
higher stresses than the static liquefaction region and encompasses a large confining
pressure range. The behaviour shows temporary liquefaction from C to D on
suppressed contraction, then strain hardening from D to E on suppressed dilation,
Appendix B Page B42

which occurs at large axial strains. Increased stability is observed with increasing
confining stress in this region.

Figure B2.14: General types of undrained effective stress paths for moist placed silty
sands sheared from a single isotropic consolidation line (Yamamuro and Lade 1997).

Yamamuro and Lade (1998) refer to the liquefaction and temporary liquefaction regions
observed at low confining stresses for very loose silty sands as indicating “reverse”
behaviour to that of clean sands. It is reverse in that, at low confining pressures,
increasing stability is observed with increasing confining pressure, eventually reaching
the earlier described “stable” region where the soil is strain hardening on shearing in
undrained loading (Figure B2.14). At confining pressures beyond this point, the
“normal” behaviour for sands is observed (i.e. temporary stability and unstable regions).
Yamamuro and Lade (1998) consider particle re-arrangement of the unstable and
highly compressive particle structure formed due to deposition under very low energy as
the mechanism for liquefaction and temporary liquefaction. At the very high pressures
associated with the instability region, they consider particle crushing to be the dominant
mechanism (Figure B2.14).
The tests by Yamamuro and Lade (1997, 1998) were undertaken from samples
prepared using the same method to the same initial void ratio, then isotropically
consolidated to the test confining pressure before undrained testing. The samples (at
different confining pressures) were tested from the same isotropic consolidation line
Appendix B Page B43

(ICL), but not from the same void ratio. Therefore, Figure B2.14 is dissimilar to Figure
B2.11 after Sladen et al (1985a) describing the collapse surface. From Figure B2.15 it
is evident that the ICL for the Nevada sand sample crosses the undrained steady state
line at two locations. Therefore, with increasing confining pressure the state parameter,
f , would go from an initially positive state to a negative state and then back to a positive
state. This could partly explain the “reverse” behaviour described by Yamamuro and
Lade (1998).
The results of testing by Yamamuro and Lade (1997) on very loose sands with trace
to some silt (percent fines less than about 10%) also highlights several other points:
• As the void ratio (or relative density) at preparation is increased, the range of
confining pressure under which static liquefaction (to zero residual strength) occurs
decreases.
• At higher confining pressures and higher density ratios the compressibility of the
sample decreases and better contact between the soil particles increases the samples
dilatant tendencies.
• At low confining pressure, the initial portion of the effective stress path for very
loose samples is relatively shallow. For these samples in the liquefaction region, the
maximum effective stress ratio at liquefaction is significantly less than for the
temporary liquefaction region (Figure B2.16).

Figure B2.15: Steady state diagram obtained from drained and undrained tests sheared
from a single isotropic consolidation line (Yamamuro and Lade 1998)
Appendix B Page B44

Figure B2.16: Reduced maximum effective stress ratio for very loose sands susceptible
to complete static liquefaction at low confining pressures (Yamamuro and Lade 1997)

Been et al (1991), based on the results of drained and undrained triaxial compression on
Erksak sand (Figure B2.17a), considered the steady state line to be a unique parameter.
For the drained tests on dense samples, Been et al (1991) comment that the samples are
generally still dilatant at the end of the test indicating the steady state has not been
reached. In addition they comment that strain localisation occurs along the surface of
rupture and therefore the void ratio is no longer uniform in the sample. They considered
the observed change in direction of the steady state line (Figure B2.17a) to be a result of
grain crushing at high confining stresses. For triaxial extension tests on the same sand
(Figure B2.17b), Been et al (1991) comment that the samples were still dilating at the
end of the test (the arrows indicating the direction of the change in mean effective
normal stress at the end of the test). Based on these results, they considered the critical
state line to be independent of the stress path.
Uthayakumar and Vaid (1998) considered that the inherent anisotropy of natural
sands has a significant effect on the strength and deformation characteristics depending
on the principal stress directions in relation to the structure of the sand. Their test
results (Figure B2.18) showed that changing the principal stress direction had a
significant effect on the stress-strain behaviour and steady state strength. As shown in
Figure B2.18, the laboratory results indicate the sample behaviour changed from strain
hardening in compression to strain softening in extension. Uthayakumar and Vaid
(1998) concluded the steady state condition is not a unique parameter for sands, but that
Appendix B Page B45

it is dependent on void ratio, the direction of loading and the intermediate principal
stress level.

Figure B2.17: Critical or steady state line of (a) drained and undrained triaxial
compression tests and (b) undrained compression and extension tests on Erksak sand
(Been et al 1991)

Ayoubian and Robertson (1998), based on the results of monotonic undrained triaxial
extension tests on Ottawa sand, concluded that the void ratio at steady state in extension
tests was affected by strain localisation. They found that void ratio redistribution was
significant from phase transformation and continued up until the failure by necking
occurred. Within the zone of necking the void ratio was significantly higher than the
average sample void ratio.
Poulos (1998) considered that for triaxial extension tests it is impossible to maintain
a uniform void ratio in the sample and that if the void ratio could be measured in the
zone of shear that it would correspond to the steady state line for triaxial compression
Appendix B Page B46

tests. The results from Ayoubian and Robertson (1998) would tend to support this.
Jefferies (1998) considered that much of the non-uniqueness reported comes from an
incorrect determination of the steady state condition. Riemer (1998), on the other hand,
indicated that the steady state line is not unique, that it is stress path dependent, as
published results of triaxial compression compared to triaxial extension, torsional shear
and simple shear indicate.
Imam et al (2002), in researching the yield surface for sands, attributed the
dependency of the undrained response of loose sand on load direction to the inherent
anisotropy associated with deposition of the sand, which existed prior to consolidation.
Higher yielding stresses are observed where the major principal stress is in the direction
of soil deposition. Therefore, for a sand sample formed by vertical deposition the
highest yielding stress in undrained loading is observed for a standard undrained triaxial
compression test, and the minimum yielding stress for a standard undrained triaxial
extension test.

Figure B2.18: Variation in effective stress path and steady state strength due to variation
in principal stress direction (Uthayakumar and Vaid 1998).
Appendix B Page B47

2.4.2 Characteristics of Sands with Fines

Byrne and Beaty (1998) raised the question of whether fines in sands were an instigator
or inhibitor of liquefaction potential. They commented that the structure of the soil had
a significant influence on the characteristics of sands with fines. Robertson and Wride
(1997) commented that from field case studies it was not clear what effect fines had,
whether the cyclic resistance of the soil is increased because of greater liquefaction
resistance or in fact the (N1)60 value of the soil decreases as a consequence of the
increase in compressibility and decrease in permeability associated with increasing fines
content.
Lade and Yamamuro (1997) undertook an extensive study on the effect of non-
plastic fines on flow liquefaction potential. They found that the flow liquefaction
potential increased with increasing fines content (Figure B2.19) and commented that the
fines had a significant role in the particle structure, resulting in highly compressible
soils and the increased flow liquefaction potential. For the Nevada 50/200 sand sample
(Figure B2.19b) the density index defining the static (or flow) liquefaction / non-
liquefaction boundary increased from 20% at zero percent fines to up to 60% at fines
contents in excess of 50%. Indicating the presence of fines generated an inherently
contractive volume tendency. Lade and Yamamuro (1997) indicate that with increasing
confining pressure the void ratio of the silty sand sample reduced due to consolidation.
This led to the movement together and better grain-to-grain contact of the larger sand
grains, resulting in increased stiffness and an increasingly dilatant tendency of the
sample.
Lade and Yamamuro (1997) recognised that the results of other researchers (Kuerbis
et al 1988; Pitman et al 1994) found that fines content increased the liquefaction
resistance of the soil. They commented that the reason for observations contrary to their
findings was due the basis of comparison used, and considered that density ratio should
be used as the basis of comparison.
Lade and Yamamuro (1997) also commented that the findings of field testing
correlations with actual earthquake induced liquefaction events indicated sands with
significant fines content were more resistant to cyclic liquefaction. They considered
that the effects of stress history and aging (creep or cementation) are likely to increase
the field resistance of silty sands. However, for newly constructed hydraulic fills and
very recent natural deposits they considered that fines content could decrease the
Appendix B Page B48

liquefaction resistance, depending on the particle structure as a result of the depositional


process.

Figure B2.19: Increased liquefaction potential with increasing non-plastic fines content
(Lade and Yamamuro 1997)

The question of the steady state being a unique parameter for cohesionless soils has also
been raised for silty sands. Yamamuro and Lade (1998) found that the steady state
condition was not unique for sands with fines when tested at different density ratios
(Figure B2.20) or when tested in drained or undrained conditions from the same density
ratio (Figure B2.15). In the latter case the steady state line was coincident above
effective confining pressures of 200 kPa, but diverged at lower confining pressures.
Yamamuro and Lade (1998) considered the divergence at low confining pressures was
due to complete static liquefaction of the undrained tests at these low effective
Appendix B Page B49

confining pressures before the tendency for dilation to a steady state condition, as
observed at higher confining pressures.
Yamamuro and Lade (1998) also observed a significant change in behaviour as a
result of strain rate effects, noting that the soil appeared to be significantly stronger at
higher strain rates. They observed that an increase in strain rate resulted in a steepening
of the effective stress path to the initial peak and an increase in the steady state strength.

Figure B2.20: Non-unique steady state lines for silty sands tested at different density
ratios (Yamamuro and Lade 1998)

2.4.3 Characteristics of Gravels and Sandy Gravels based on Laboratory Testing

Evans and Zhou (1995) carried out cyclic undrained triaxial tests on gap-graded
mixtures of medium sand and fine to medium gravel, tested with varying gravel
contents. They found that the cyclic liquefaction resistance (based on a 5% double
amplitude strain liquefaction criterion) increased with increasing gravel content.
Dawson (1994), as part of his research into liquefaction flow slides of coal mine
waste dumps in the Rocky Mountains, undertook a series of undrained triaxial
compression tests on well-graded sandy gravels. Samples were prepared by moist
placement and then isotropically consolidated to pressures ranging from 50 to 500 kPa.
Figure B2.21 presents the results from the Fording sandy gravel sample. Dawson
(1994) found that at low confining pressures the samples were unstable and contractive,
and therefore in undrained loading the steady state strength was less than the peak
strength. At increasing confining pressures he found that the samples became more
Appendix B Page B50

stable (i.e. temporary instability condition observed). The consolidation lines were
roughly parallel to the steady state line (Figure B2.21b) and the brittleness index, Ib
( I b = ∆q / qmax where q is the deviatoric stress), decreased with decreasing void ratio to
a brittleness index of 0% at void ratios less than or equal to 0.3.

Figure B2.21: Fording sandy gravel sample (a) undrained triaxial compression test
results and (b) consolidation and steady states (Dawson et al 1998)

2.5 RESIDUAL UNDRAINED SHEAR STRENGTH OF FLOW LIQUEFIED SOILS

Methods developed for assessment of the residual undrained shear strength of liquefied
soils include:
Appendix B Page B51

• Empirical correlation based on back analysis of mostly earthquake induced


liquefaction flow failures (Stark and Mesri 1992; Seed 1987; Seed and Harder 1990;
Baziar and Dobry 1995). These correlations are presented in Section 3.3.5 of
Chapter 3.
• Laboratory test methods, from which theories were developed to explain the
observed behaviour of sands (Ishihara (1993) amongst others).

From review of the published literature it is apparent that there is a large uncertainty
in the residual undrained strength determined by these methods and that the results from
the literature, particularly from laboratory testing, present conflicting results.
Castro (1998) commented that different back-analyses of similar case histories have
led to substantially different interpretations. He considered the strong reliance in
practice on these empirical correlations from case histories to be unwarranted. Castro
(1998) also commented that drainage (of the liquefied soil) affects the strength and
therefore travel of the liquefied soil. He recommended that the likelihood or not of
drainage should be assessed on a case by case basis taking into account the size of the
failure zone, proximity to drainage boundaries, soil permeability and the duration of the
post failure movement.
Martin (1998) considered that the undrained, large strain shear strength (residual
undrained shear strength) determined from laboratory testing is not necessarily the
strength mobilised under field conditions due to issues relating to pore water pressure
redistribution, void ratio changes, effective stress path to failure, and the influence of
interface conditions associated with stratified silty and sandy soils impacting the
residual strength mobilised in the field.
There is significant debate over the merits of laboratory testing as a method of
evaluating the undrained residual shear strength of liquefied soils. At the National
Science Foundation workshop on the shear strength of liquefied soils (NSF 1998) no
consensus could be reached on the type of laboratory testing to be undertaken due to the
observed (and sometimes contradictory) behaviour in laboratory tests. The main areas
of debate related to; the uniqueness or otherwise of the critical or steady state line
(discussed in Sections 2.4.1 and 2.4.2), interpretation of the laboratory results, the
effects of strain controlled or load controlled testing, and the limitations of strain in
laboratory testing.
Appendix B Page B52

A significant number of laboratory triaxial undrained test results show what is


referred to as a “quasi steady state”, where post peak the deviatoric stress initially
decreases to the quasi steady state and then with further straining beyond the quasi
steady state increases (Figure B2.9a). The results from Ishihara (1993) indicate that the
quasi steady state condition is non-unique in that it is dependent on the initial fabric of
the sample tested (refer Section 2.4.1). Kramer (1998) and Gutierrez (1998) comment
that given the quasi steady state is not a unique state it cannot be considered a
fundamental state on which constitutive models can be based. Jefferies (1998)
comments that the strength gain from post peak dilation should not be relied upon and
that whilst research into strain localisation continues the quasi steady state should be
used as a lower bound estimate of the post liquefaction strength.
Yamamuro and Lade (1998) report that strain rate does have an effect on the steady
state condition of silty sands. Conversely, Been et al (1991) and Poulos (1998)
comment that strain rate has no effect. Therefore, the results are inconclusive with
respect to strain rate effects on the steady state condition.
De Alba (1998) questions the validity of the triaxial test for prediction of post
liquefaction behaviour due to the strain limitation of these cells in comparison to the
high strains observed in the field. He commented that triaxial samples could not be
deformed beyond 20 to 25% axial strain without stress calculations becoming
meaningless due to sample distortion. Verdugo et al (1995) commented, from the
results of their triaxial testing, that natural non-homogeneous sample structure couldn’t
be broken down even at large deformations in the triaxial cell.
APPENDIX C

Embankments on Soft Ground


Appendix C Page i

TABLE OF CONTENTS

1.0 ANALYSIS OF EXCESS PORE WATER PRESSURE


RESPONSE AND PARTIAL DRAINAGE DURING THE
INITIAL STAGES OF EMBANKMENT
CONSTRUCTION ...................................................................... C1
1.1 Summary of the Literature ............................................................................. C1
1.2 Analysis of the Pore Water Pressure Response from the Case Study
Data ................................................................................................................ C2

2.0 CASE STUDIES OF POST CONSTRUCTION


DEFORMATION BEHAVIOUR ............................................... C8
2.1 Muar Test Embankments................................................................................ C8
2.2 Olga-C Test Embankment ............................................................................ C17
2.3 Queensborough Test Embankment .............................................................. C21

3.0 CASE STUDY EXCEPTIONS TO GOOD INDICATORS


OF AN IMPENDING FAILURE.............................................. C25
3.1 Excess Pore Water Pressure Response......................................................... C25
3.2 Pre-Failure Deformation Response .............................................................. C28

4.0 POST-FAILURE DEFORMATION ANALYSIS.................... C30


4.1 Rio Test Embankment, Rio de Janiero......................................................... C31
4.2 Muar-F Test Embankment, Malaysia........................................................... C33
4.3 St. Alban-A Test Embankment, Quebec, Canada ........................................ C34
4.4 Portsmouth Test Embankment, New Hampshire, USA ............................... C36
Appendix C Page ii

LIST OF TABLES
Table C1.1: Field results of pore water pressure response during initial loading.......... C4

LIST OF FIGURES
Figure C1.1: Field profiles of the rate of pore water pressure development during
initial stages of construction.................................................................................... C3
Figure C1.2: Field profiles of the rate of pore water pressure development during
the initial stages of construction. ............................................................................. C7
Figure C2.1: Section of the Muar-EC test embankment and interpreted surface of
rupture (MHA 1989a) ........................................................................................... C10
Figure C2.2: Contours of normalised shear-stress ratio at a fill height of 4 m from
finite element modelling of the Muar-F test embankment (Indraratna et al
1992) ..................................................................................................................... C11
Figure C2.3: Vertical inclination angle profiles with increasing fill height at
Inclinometer I1, located at the toe of Muar-EC test embankment. ....................... C11
Figure C2.4: Muar-EC, excess pore water pressure response with time for
piezometers (a) under the centre of the embankment, and (b) under the toe of
the embankment. ................................................................................................... C12
Figure C2.5: Muar-EC, settlement and maximum lateral displacement with time...... C13
Figure C2.6: Section of the Muar-3C test embankment (MHA 1989b)....................... C14
Figure C2.7: Muar-3C, excess pore water pressure response with time for
piezometers (a) under the centre of the embankment, and (b) under the toe of
the embankment. ................................................................................................... C15
Figure C2.8: Vertical inclination angle profiles with increasing fill height at
Inclinometer I1, located at the toe of Muar-3C test embankment......................... C16
Figure C2.9: Muar-3C, settlement with time under the embankment.......................... C16
Figure C2.10: Muar-3C, lateral displacement at the toe of the embankment with
time........................................................................................................................ C17
Figure C2.11: Section A of the Olga-C test embankment with no wick drains
(Lavallée et al 1992).............................................................................................. C18
Figure C2.12: Olga-C, excess pore water pressure response with time at Section A
(a) below the centre line, (b) below the upper slope and (c) below the berm. ...... C19
Figure C2.13: Olga-C, vertical inclination angle profiles at various times at
Inclinometer INC1, located at about mid berm, at Section A. .............................. C20
Appendix C Page iii

Figure C2.14: Queensborough test embankment, section of test embankment


showing instrumentation (Jardine and Hight 1987). ............................................. C22
Figure C2.15: Queensborough test embankment, excess pore water pressure
response versus (a) applied embankment load and (b) time. ................................ C23
Figure C2.16: Settlement at the centre and maximum lateral displacement at the
toe of the Queensborough test embankment. ........................................................ C24
Figure C3.1: Interpretation of effective stress path for heavily over-consolidated
clays under embankments (Leroueil et al 1978a).................................................. C26
Figure C3.2: Effective stress paths and pore water pressure generation associated
with different types of yield conditions (Leroueil and Tavenas 1986). ................ C27
Figure C4.1: Pre and post-failure geometry of Rio test embankment (Ramalho-
Ortigao et al 1983b) .............................................................................................. C32
Figure C4.2: Pre and post-failure geometry of Muar-F test embankment (Brand
and Premchitt 1989) .............................................................................................. C33
Figure C4.3: Pre and post-failure geometry of St. Alban-A test embankment (La
Rochelle et al 1974) .............................................................................................. C35
Figure C4.4: Pre and post-failure geometry of Portsmouth test embankment (Ladd
1972) ..................................................................................................................... C36
Appendix C Page C1

1.0 ANALYSIS OF EXCESS PORE WATER PRESSURE


RESPONSE AND PARTIAL DRAINAGE DURING THE
INITIAL STAGES OF EMBANKMENT CONSTRUCTION

1.1 SUMMARY OF THE LITERATURE


Based on the results of field observations, Leroueil et al (1978a) and Leroueil and
Tavenas (1986) consider that the process of consolidation and therefore partial drainage
occurs during the initial stages of embankment construction whilst the clay foundation
is in an over-consolidated condition. The assumption of undrained conditions can
therefore be erroneous in estimation of pore water pressure response and deformation
behaviour.
Leroueil et al (1978a) evaluated the excess pore water pressure response from
approximately 95 piezometers located under the embankment centreline of a total of 30
embankments constructed over soft ground. Only the piezometers under the
embankment centreline were evaluated because, at this location, the stresses can be
estimated with reasonable accuracy and it can be assumed that there is no rotation of the
principal stress axes. Leroueil et al (1978a) used Skempton’s compounded pore water
pressure coefficient, B ( B = ∆ u ∆σ 1 , refer to Equation 4.2 in Section 4.2.1 of Chapter
4 for a full definition), to evaluate the excess pore water pressure response. The results
of their analyses (Figure 4.3 in Section 4.2.1 of Chapter 4) indicate that during the initial
stages of construction, whilst the clay foundation is in an over-consolidated condition,
the shape of the excess pore water pressure profile with depth is similar to that of a
consolidation isochrone. The maximum B1 value (the subscript 1 referring to the initial
stage of loading whilst the clay foundation is in an over-consolidated condition) is
located in the central portion of the soft clay layer and reduces to approximately 0 at the
drainage boundaries.
Leroueil and Tavenas (1986) considered that a lack of saturation could also result in
the observation of pore water pressure generation smaller than expected for undrained
conditions. However, the isochronal shape of the pore water pressure profile and the
higher excess pore water pressures developed with increasing rate of construction
observed at the St. Alban test embankments indicated that consolidation was significant.
Appendix C Page C2

Leroueil and Tavenas (1986) commented that the degree of consolidation varies
from case to case, depending on the coefficient of consolidation, cv, of the over-
consolidated clay, the length of the drainage path and the rate of loading. Crooks
(1987) considers that the stiffness and permeability of the foundation, rate of loading
and drainage boundary conditions affect the degree of consolidation during initial
loading. He comments that clay foundations with high stiffness (in an over-
consolidated condition) would be expected to generate only moderate excess pore water
pressures during initial loading, whilst for clays of lower stiffness development of a
more significant excess pore water pressure response would be expected. Based on the
comments by Crooks (1987) it could be expected therefore that partially cemented,
sensitive clay foundations, which have relatively high stiffness on initial loading, would
be expected to show a more significant reduction in excess pore water pressure
generation than high plasticity clays of low sensitivity, which are typically of
significantly lower stiffness.
Tavenas and Leroueil (1980) proposed an equation of best fit (Equation C1.1) to the
data from their analysis (Figure 4.3 in Section 4.2.1 of Chapter 4) as a means of
predicting the initial rate of pore water pressure development.

B1 = 0.6 − 2.4(z D − 0.5 )


2
(C1.1)

where z is the depth of the point at which B1 is to be estimated and D is the thickness of
the soft clay deposit.

1.2 ANALYSIS OF THE P ORE WATER P RESSURE R ESPONSE FROM THE


CASE S TUDY DATA
The purpose of this section of Appendix C is to evaluate the influence of the properties
of the soft clay foundation and the rate of loading on the pore water pressure response of
piezometers located under the embankment centreline during the initial stages of
construction prior to yielding.
The excess pore water pressure responses from a total of 55 piezometers, located
within the clay foundation under the embankment centreline, from 16 case studies have
been analysed. For each piezometer the average B1 value has been determined from the
pore water pressure response during construction whilst the foundation at that location
Appendix C Page C3

has been assessed to be in an over-consolidated condition (i.e. prior to yielding). The


results of the analysis are presented in Table C1.1. Plots of the profiles of initial rate of
pore water pressure development ( B1 ) versus the normalised depth (z/b) are presented
in Figure C1.1, where z is the depth below ground surface level and b is the half width
of the embankment modelled as a rectangular strip load.

King's Lynn
James Bay
0.2 Portsmouth
St. Alban-A
St Alban-B
Muar-F
0.4 Muar-EC
Muar-3C
Rio
Normalised Depth (z/b)

Cubzac-A
0.6 Kalix
Thames
Sackville
Queensborough
0.8 Olga-C
mean total stress ratio

Notes:
1 The mean total stress ratio is
based on the undrained
response of an infinite strip load
on an elastic half space.

1.2 z = depth below ground surface


level
b = half width of the embankment

1.4
0 0.2 0.4 0.6 0.8 1
B 1 value, and Mean Total Stress Ratio

Figure C1.1: Field profiles of the rate of pore water pressure development during initial
stages of construction.
Appendix C Page C4

Table C1.1: Field results of pore water pressure response during initial loading
Case Study Piezometer Rate of B 1 = ∆u / ∆σ v OCR Soil Description
1
No. Depth, z/b * Construction Field Undrained
z (m) (m/day) Response*2
King’s Lynn 214 3.3 0.44 0.67 0.07 0.73 2.1 CH marine clay with
(Wilkes 1972) 215 4.7 0.63 0.07 0.64 1.8 peat layers.
James Bay A-1 1.5 0.07 0.15 0.29 0.96 14 CL/ML leached marine
(Dascal & A-2 4.6 0.2 0.58 0.88 3.3 clay, very sensitive,
Tounier 1975) A-3 7.6 0.33 0.64 0.79 2.0 brittle.
A-4 10.7 0.47 0.34 0.72 1.7
Portsmouth V-5 5.8 0.28 0.19 0.62 0.83 1.25 CL leached marine
(Ladd 1972) V-9 7.3 0.36 0.61 0.78 1.25 clay, sensitive, brittle.
V-10 10.9 0.53 0.44 0.69 1.2
Mastemyr*3 P1 3.5 .41 0.19 0.87 0.63 2.5 CL/ML leached marine
(Clausen 1972) P2 6.0 .70 0.60 0.44 1.2 clay, very sensitive,
P3 11.0 1.28 0.31 0.21 1.1 brittle.
St. Alban-A D1 1.5 0.18 0.45 0.07 0.89 2.4 CL/ML leached marine
(La Rochelle et D2 4.6 0.55 0.68 0.68 1.7 clay, cemented,
al 1974) D3 9.3 1.12 0.40 0.47 1.6 sensitive, brittle.
C1 3.0 0.36 0.20 0.78 2.0
St. Alban-B T1 2.58 0.29 0.19 0.35 0.83 2.1 CL/ML leached marine
(Leroueil et al T2 5.0 0.57 0.28 0.67 1.7 clay, cemented,
1978b) G3 9.3 1.06 0.21 0.48 1.6 sensitive, brittle.
T4 12.0 1.37 0.08 0.41
Muar-F P1 0.8 .0.5 0.054 0.18 0.97 >10 CH marine clay, low
(Brand & P2 4.5 0.27 0.22 0.83 1.6 sensitivity, ductile.
Premchitt 1989) P3 9.0 0.55 0.60 0.68 1.4
P4 13.5 0.82 0.61 0.56 1.25
Muar-EC P4 4.6 0.22 0.017 0.37 0.87 1.6 CH marine clay, low
(MHA 1989a) P5 9.0 0.44 0.58 0.73 1.3 sensitivity, ductile.
P6 13.5 0.66 0.48 0.63 1.3
Muar-3C P4 4.6 0.35 0.05 0.26 0.78 1.6 CH marine clay, low
(MHA 1989b) P5 9.1 0.70 0.20 0.62 1.3 sensitivity, ductile.
P6 13.6 1.05 0.54 0.49 1.3
Rio de Janiero 1 2.0 0.17 0.10 0.27 0.90 3.0 CH marine clay, low
(Ramalho- 2 3.0 0.26 0.27 0.84 2.4 sensitivity, ductile.
Ortigao et al 3 4.0 0.35 0.33 0.79 2.0
1983a) 4 5.0 0.43 0.68 0.74 1.9
5 8.0 0.69 0.69 0.62 1.6
Cubzac-A 112 4.0 0.32 0.12 to 0.75 0.38 0.81 1.6 CH clay.
(Blondeau et al 132 6.0 0.48 0.63 0.72 1.3
1977)
Kalix P1 2.5 0.33 0.35 0.52 0.79 2.2 MH/OH/CH marine
(Holtz & Holm P2 5.0 0.67 0.61 0.62 1.6 silt/clay, sensitive,
1979) P3 7.5 1.0 0.57 0.50 1.4 brittle.
Thames C1 2.3 0.22 0.10 0.18 0.87 2.3 CL/ML marine clay
(Marsland & C2 4.3 0.42 0.24 0.74 1.8 with peat layers.
Powell 1977) C3 5.7 0.55 0.31 0.68 1.7
C4 7.4 0.72 0.35 0.61 1.4
C5 10.3 1.0 0.29 0.50 1.4
C6 11.7 1.14 0.25 0.46 1.3
*1 b is the half width of the embankment modelled as rectangular strip load.
*2 The undrained response is the estimated pore water pressure response assuming an infinite strip load
on an elastic half space.
3
* Mastemyr is the only circular embankment of the data set. It is not included in Figure C1.1.
Appendix C Page C5

Table C1.1: Field results of pore water pressure response during initial loading
(continued)
Case Study Piezometer Rate of B 1 = ∆u / ∆σ v OCR Soil Description
1
No. Depth, z/b * Construction Field Undrained
z (m) (m/day) Response*2
Sackville 15 2.0 0.21 0.23 0.32 0.87 1.6 ML/MH intertidal salt
(Rowe et al 16 4.0 0.42 0.53 0.74 1.1 marsh deposit.
1995) 17 6.0 0.63 0.44 0.64 1.1
Queensborough P5/2 7.6 0.38 0.07 0.67 0.77 1.1 Alluvial silty clays of
(Jardine & Thames estuary, UK
Hight 1987) (similar to Thames
case)
Olga-C A108 2.0 0.07 0.15 0.23 0.96 12 CH varved clay
(Lavallée et al A107 4.5 0.16 0.44 0.91 4.2
1992) A106 7.5 0.26 0.74 0.84 2.5
A105 10.5 0.36 0.71 0.78 1.8

Muir Wood (1990) considered that by assuming undrained conditions during loading
and that the foundation behaviour is elastic prior to yield (i.e. when in an over-
consolidated condition) the excess pore water pressure response could be derived
directly from the change in mean total stress. In Table C1.1 comparative values of the
estimated rate of pore water pressure development ( B1 ) assuming an undrained
response of an infinite strip load on elastic half space are given for each piezometer. In
Figure C1.1 the plot of mean total stress ratio for undrained conditions is also given for
comparative purposes. Figure C1.2 presents the same results with B1 normalised with
respect to the mean stress ratio, thus giving an estimate of the percent drainage, versus
the depth, z, below ground surface level.
In comparison to the undrained profile for an infinite strip on elastic half space, the
excess pore water pressure profile for most case studies is typical of a consolidation
isochrone for upward drainage only (Figure C1.2). The B1 profiles of four case studies
(Portsmouth, James Bay and St. Alban A and B), all having foundations comprising low
plasticity, sensitive, leached marine clays, show signs of drainage toward a second
drainage boundary at depth. Therefore, the field observations confirm that the length of
the drainage path is a significant factor on the degree of consolidation.
However, apart from this, the results show a high degree of variability from case to
case, with no clear correlation with respect to rate of loading between different sites,
material type (from very low to high sensitive and low to high plasticity clays) or likely
material stiffness. It is considered that these observations reflect that permeability is a
key factor in controlling the variation of cv between case studies and therefore the
Appendix C Page C6

degree of consolidation that occurs during the initial stages of loading. The
observations that indicate permeability is a significant factor are as follows:
• The pore water pressure profiles for the Muar test embankments (Figure C1.1 and
Figure C1.2) indicate a significant increase in the value of B1 at greater than about 9
m depth, which is consistent with depth of the interface between the upper and
lower clay strata. Brand and Premchitt (1989) indicate that the permeability of the
upper clay strata is about 4 times greater than that of the lower layer.
• In the case of King’s Lynn it is considered that the relatively high permeability of
the peat layer has resulted in the low B1 values calculated. Piezometers 214 and
215 are located slightly above and within the extensive peat layer respectively. The
permeability of the peat layer is reported by Wilkes (1972) to be 10 to 15 times
greater than the permeability of the CH alluvial clays above and below the peat.
• In the case of the Rio test embankment the pore water pressure profile would
suggest a lower permeability at depth in the soft clay stratum as indicated by the
abrupt increase in B1 between Piezometer 3 (4 m depth) and Piezometer 4 (5 m
depth).
• The lack of correlation from case to case to the rate of construction.
• No correlation to material stiffness. For near normally consolidated soils, the low
plasticity, sensitive, brittle leached marine clays (some of which are cemented)
would be expected to have a relatively high stiffness in comparison to high plasticity
clays low sensitivity. The pore water pressure profiles show no distinction between
these material types.

For sites analysed with more than one test embankment, the St. Alban test
embankments (Leroueil et al 1978a) do indicate a correlation with respect to rate of
loading. St. Alban A was constructed at a rate approximately 2 to 2.5 times greater than
the St. Alban B test embankment and the B1 profile for St. Alban A is significantly
greater than for St. Alban B, as expected. However, for the Muar tests embankments
(Muar-F, Muar-EC and Muar-3C) no correlation is evident with respect to rate of
loading.
In 7 of the 16 case studies analysed the rate of pore water pressure generation at a
depth below 4 to 6 m (Figure C1.2) exceeded 80% of the pore water pressure profile
assuming undrained conditions. This indicates that near undrained conditions are
Appendix C Page C7

observed at depth in a significant proportion of cases during the initial stages of


construction whilst the foundation is in an over-consolidated condition.

King's Lynn
James Bay
4
Portsmouth
St. Alban-A
St Alban-B
6 Muar-F
Muar-EC
Depth (m)

Muar-3C
8 Rio
Cubzac-A
Kalix
10 Thames
Sackville
Queensborough
12 Olga-C
Mastemyr

14

16
0 0.25 0.5 0.75 1 1.25 1.5
(B1 ) / (Mean Total Stress Ratio)
Figure C1.2: Field profiles of the rate of pore water pressure development during the
initial stages of construction.

In the case of the Mastemyr test embankment (leached marine deposit of very soft,
quick, low plasticity clay) the observed field pore water pressure response stands out as
having a very high rate of pore water pressure generation during the initial stages of
construction, greater than that estimated assuming undrained conditions. It is not
certain of why this behaviour is observed at Mastemyr or why the rate of pore water
pressure generation is much greater than at the other test embankment sites. It may
possibly be related to a very low permeability of the structured foundation or to its
Appendix C Page C8

stress-strain relationship. Leroueil et al (1978b) could not establish the reasons for the
pore water pressure response at this site.
The observation of a pore water pressure response greater than the estimate for
undrained conditions is likely to be associated with the assumption of the foundation as
a half space and that the foundation behaves elastically prior to yielding for the
undrained case considered.

2.0 CASE STUDIES OF POST CONSTRUCTION


DEFORMATION BEHAVIOUR

2.1 M UAR TEST EMBANKMENTS


The Malaysian Highway Authority (MHA) undertook a major study of embankment
construction on soft Malaysian marine clays. The study comprised the construction of
13 well-monitored embankments for assessment of different methods of ground
improvement (MHA 1989a) and one embankment constructed to failure (Brand and
Premchitt 1989). Three of the embankments have been studied in detail as part of this
research project, they are briefly summarised as follows:
• Muar-F – the test embankment constructed to failure (Brand and Premchitt 1989).
The fill height at failure was 5.4 m.
• Muar-EC – Scheme 6/1 (MHA 1989a) comprising chemical injection of the
foundation to 7.6 m depth prior to embankment construction. The embankment
failed at a fill height of 4.68 m.
• Muar-3C – 3 m high control embankment on an untreated foundation. The
embankment was constructed to a height of 3.9 m, and no failure occurred.

All three embankments were constructed with slopes of 2H to 1V (horizontal to


vertical), but with crest widths ranging from 20 m to 34 m. Highly plastic residual
granite was used in all cases as the embankment material. It was placed in thin layers
and well compacted using rollers. One noted variation was that the embankment filling
for Muar-F appears to have been compacted slightly dry of optimum moisture content
compared to slightly wet of optimum for the Muar-EC and Muar-3C embankments.
Compacting on the dry side of optimum would be expected to result in a lower
compressibility and higher undrained strength of the fill.
Appendix C Page C9

The rate of filling also varied between the case studies analysed. Muar-F was filled
at a steady rate of 0.054 m/day over a period of 100 days. Muar-3C and Muar-EC were
constructed in multiple stages ranging from 0.2 m to 1.8 m stage heights with rest
periods of 10 to 100 days between construction periods, with a total construction time of
290 and 270 days respectively. The overall average rate of construction of both these
embankments was less than 0.02 m/day.
Details of the subsurface conditions and depositional history of the marine clays at
Muar are discussed by MHA (1989a). In summary, the upper 17 to 18 m of the
subsurface comprises recent marine clay deposits, with the deposition considered to
have occurred in two sequences. A lower sequence of very soft silty clay with
considerable shells of approximately 9 m thickness deposited about 5000 years B.P.
(before present); and an upper sequence of very soft clay, of approximately 8 m
thickness, where deposition commenced some 3000 to 4000 years B.P. A discontinuous
layer of sandy silty clay separates the upper and lower clay sequences. A weathered
crust of approximately 2 m thickness has since developed in the upper clay profile,
largely as a result of the drop in sea level.
The geotechnical properties of the subsurface layers are summarised as follows:
• Weathered crust of heavily over-consolidated (OCR maximum of about 10),
fissured, high plasticity clay.
• Upper layer of very soft, high plasticity silty clays (undrained shear strength of 10 to
17.5 kPa measured by vane, liquid limit of 65 to 120 % and plastic limit of 26 to 63
%). The clay is lightly over-consolidated (OCR decreasing from 2.0 to 1.3) and has
an average liquidity index of 1.3 to 1.5.
• Lower layer of soft, high plasticity silty clay with considerable shells (undrained
shear strength of 19 to 30 kPa measured by vane). The clay is lightly over-
consolidated (OCR of 1.2 to 1.3) and has an average liquidity index of 1.25 to 1.5.

Remoulded vane tests indicate a remoulded strength ranging from 2 kPa at 2.6 m
depth to 7 kPa at 17 m depth for the soft clays, correlating to an average sensitivity
index of 4. Estimation of the permeability from laboratory oedometer tests indicates the
upper clay layer to be of higher permeability than the lower clay layer, approximately
4 × 10 −9 m/s and 10 −9 m/s respectively.
Appendix C Page C10

The deformation behaviour and excess pore water pressure response for the Muar-
EC and Muar-3C embankments, both during and post construction, is discussed in the
following sub-sections.

2.1.1 Muar-EC Test Embankment

Figure C2.1 is a section of the Muar-EC test embankment showing the position of
instrumentation together with the interpreted surface of rupture and subsurface profile.
Indraratna et al (1992) undertook finite element modelling of the Muar-F test
embankment. The analysis showed that a zone of high shear stress concentration
developed at a depth of 7.5 to 8 m below the slope of the embankment at embankment
heights significantly below the reported failure thickness. Figure C2.2 shows the shear
stress concentration for a fill height of 4 m (1.4 m less than the failure height of 5.4 m).
The plot of vertical inclination angle profiles for Inclinometer I1 (Figure C2.3),
located at the toe of the Muar-EC embankment, clearly shows the development of a
localised failure zone at 8 to 9.5 m depth when the embankment height was at 4.68 m.
The eventual surface of rupture was located within this localised failure zone, as was
Piezometer P2.

Figure C2.1: Section of the Muar-EC test embankment and interpreted surface of
rupture (MHA 1989a)
Appendix C Page C11

Figure C2.2: Contours of normalised shear-stress ratio at a fill height of 4 m from finite
element modelling of the Muar-F test embankment (Indraratna et al 1992)

6
Depth (m)

8
localised failure
zone
10

12
Fill height = 0.8 m
14
Fill height = 2.38 m

Fill height = 4.18 m


16
Fill height =4.40 m
18 Fill height = 4.68 m

20
-10 -5 0 5 10 15 20
Vertical Inclination Angle (rad x 102)

Figure C2.3: Vertical inclination angle profiles with increasing fill height at
Inclinometer I1, located at the toe of Muar-EC test embankment.

The excess pore water pressure response with time of Piezometer P2 (Figure C2.4b)
shows a significant increase in excess pore water pressure after the end of construction
Appendix C Page C12

(Day 269) leading up to the eventual failure condition at Day 300, some 30 days after
completion of construction. The response is significantly different to that of the other
piezometers under the centre of the embankment and also at 5.3 m and 14.3 m depth
below the toe of the embankment, which remain relatively steady in comparison over
the same time period. It is considered that the pore water pressure response at P2 is
associated with shear induced pore water pressures within the localized failure zone
leading up to the embankment failure.

a) 120 Failure at 300 days 5


P4 (centre, 4.6 m)
4.5
Excess Pore Water Pressure (kPa)

P5 (centre, 9 m)
100
P6 (centre, 13.5 m) 4
Fill Height (m)
3.5
80

Fill Height (m)


3

60 2.5

2
40
1.5

1
20
0.5

0 0
0 50 100 150 200 250 300 350
Time (Days)

b) 70 Failure at 300 days 5


P1 (toe, 5.3 m)
4.5
60 P2 (toe, 9 m)
Excess Pore Water Pressure

P3 (toe, 14.3 m) 4

50 Fill Height (m)


3.5
Fill Height (m)

3
40
(kPa)

2.5
30
2

20 1.5

1
10
0.5

0 0
0 50 100 150 200 250 300 350
Time (days)

Figure C2.4: Muar-EC, excess pore water pressure response with time for piezometers
(a) under the centre of the embankment, and (b) under the toe of the embankment.
Appendix C Page C13

Comparison of the settlement behaviour at Point S4 (within the embankment failure


zone) with Point S5 (outside the failure zone) in Figure C2.5 shows the increasing
settlement at S4 above that of S5 starting from fill heights greater than about 4.2 m
(0.48 m below the failure height) and particularly evident after the end of construction.
The maximum lateral displacement in Inclinometer I1 at the embankment toe continued
to increase at a relatively steady rate from the end of construction up to the embankment
failure (Figure C2.5). It is considered that the lateral displacement at the end of
construction is predominantly associated with plastic deformation within the localised
failure zone between 8 and 9.5 m depth and is reflected in the increasing settlement
observed at S4 above that observed at S5.

-200 5

0 4.5
Failure on
Day 300 4
200
Settlement and Lateral

3.5
Displacement (mm)

400

Fill Height (m)


3
600
2.5
800
2
1000
1.5
1200 Settlement, S4 (centre to crest) 1
Settlement, S5 (centre)
1400 Displacement at toe, I1 (3.8m depth) 0.5
Fill Height (m)
1600 0
0 50 100 150 200 250 300 350 400 450
Time (Days)
Figure C2.5: Muar-EC, settlement and maximum lateral displacement with time.

2.1.2 Muar-3C Test Embankment

Figure C2.6 is a section of the Muar-3C test embankment showing the position of
instrumentation and the interpreted subsurface profile.
The excess pore water pressure response of Piezometer P2, located at 9 m depth
below the embankment toe, continues to increase for a period of 110 to 120 days after
the end of construction (Figure C2.7). Piezometer P5, located at 9 m depth below the
centre of the embankment, shows a similar increase, albeit to a lesser extent, whilst the
remainder of the piezometers are relatively steady. It is considered that the increasing
Appendix C Page C14

pore water pressure is associated with the relatively high level of shear stress in the
vicinity of Piezometer P2.
The rise in excess pore water pressure in Piezometer P2 at Muar-3C is at an average
of 0.065 kPa/day for 120 days (Day 290 to 410). In comparison, the rise in excess pore
water pressure in Piezometer 2 at the Muar-EC embankment was on average 0.62
kPa/day for the 30 days leading up to the failure, a magnitude of 10 times greater than at
Muar-3C.

Figure C2.6: Section of the Muar-3C test embankment (MHA 1989b)

a) 60 4.5
End of construction
4
Excess Pore Water Pressure (kPa)

50
3.5

40 3
Fill Height (m)

2.5
30
2

20 1.5
P4 (centre, 4.6 m)
P6 (centre, 13.6 m) 1
10 P5 (centre, 9.1 m)
Fill Height (m) 0.5

0 0
0 100 200 300 400 500 600
Time (Days)
Appendix C Page C15

b) 60 4.5
End of construction
4
Excess Pore Water Pressure (kPa)
50
P1 (toe, 4.6 m)
3.5
P2 (toe, 9 m)
P3 (toe, 13.5 m)
40 3

Fill Height (m)


Fill Height (m)
2.5
30
2

20 1.5

1
10
0.5

0 0
0 50 100 150 200 250 300 350 400 450 500 550
Time (Days)
Figure C2.7: Muar-3C, excess pore water pressure response with time for piezometers
(a) under the centre of the embankment, and (b) under the toe of the embankment.

The plot of vertical inclination angle of Inclinometer I1 at the toe of the Muar-3C
embankment (Figure C2.8) compared with the Muar-EC embankment (Figure C2.3)
indicates that a localised failure condition has not been reached at depth under the toe of
the embankment at Muar-3C. It is concluded therefore, that the post construction
increase in excess pore water pressure at the Muar-3C embankment is due to yielding of
the normally consolidated soft clays and the strain rate dependent reduction in the limit
state surface (refer to the discussion in Section 4.4.2 of Chapter 4 and further discussion
below).
With respect to the deformation behaviour of Muar-3C it is observed that:
• Large settlements (Figure C2.9) occur during periods of steady or increasing excess
pore water pressures under the centreline of the embankment.
• A reduction in the rate of lateral displacement (Figure C2.10) after construction
occurs at about Day 420, and is coincident with the time that excess pore water
pressures started to decrease in Piezometer P2.
Appendix C Page C16

6 Fill Height = 1.8 m (Day 17)


Fill Height = 2.7 m (Day 56)
8
Depth (m)

Fill Height = 2.7 m (Day 140)

10 Fill Height = 3.56 m (Day 164)


Fill Height = 3.9 m (Day 322)
12 Fill Height = 3.9 m (Day 518)

14

16

18

20
-10 -5 0 5 10
2
Vertical Inclination Angle (rad x 10 )

Figure C2.8: Vertical inclination angle profiles with increasing fill height at
Inclinometer I1, located at the toe of Muar-3C test embankment.

0 5

4.5
Settlement (mm), Heave is -ive

200 4

3.5 Fill Height (m)


400 S5 (centre) 3
S4 (centre to crest)
Fill Height (m) 2.5

600 2

1.5

800 1

0.5

1000 0
0 100 200 300 400 500 600
Time (Days)
Figure C2.9: Muar-3C, settlement with time under the embankment.
Appendix C Page C17

250 5

End of construction
Lateral Displacement at toe (mm)
200 4

Fill Height (m)


I1 @ 3.7 m depth (toe)
150 3
I1 @ surface (toe)
Fill Height (m)

100 2

50 1

0 0
0 100 200 300 400 500 600
Time (Days)
Figure C2.10: Muar-3C, lateral displacement at the toe of the embankment with time.

It is concluded that the pore water pressure response and deformation behaviour indicate
that the processes of consolidation and undrained creep (as a result of the strain rate
dependence of the limit state surface) are occurring concurrently. Consolidation alone
would result in deformations in the soft clay foundation as pore water pressures
dissipate. However, the observations of the excess pore water pressure response (Figure
C2.7) indicate that the pore water pressure is either relatively steady or else increases
over most of the post construction monitoring period (Day 290 to 520), during which
period large settlements are measured.
It is considered that the pore water pressure response is due to the strain rate
dependency of the limit state surface (refer Section 4.4.2 of Chapter 4). As the limit
state surface contracts (assuming for the moment it is not expanding due to pore water
dissipation), the effective stress state of the normally consolidated soil on the yield
surface must also contract resulting in the increase in pore water pressure. It is in effect
an undrained “creep” response.

2.2 OLGA-C TEST EMBANKMENT


The Olga-C test embankment (Figure C2.11) was constructed in 1990 at the NBR
Complex site at Matagami, Quebec, Canada. The purpose of the Olga-C test
embankment was to assess the performance of wick drains.
Appendix C Page C18

The embankment was divided into four sections, three with various wick drain
configurations in the foundation and one with no wick drains. The section without wick
drains, Section A of the Olga-C test embankment, has been analysed as part of this
research.
In the early 1970s the test embankment Olga-A was constructed to failure (located
very close to the Olga-C site). The failure occurred at an embankment height of about
half that predicted based on a shear strength profile determined from field vane testing.
The subsurface conditions at the Olga C test site (Lavallée et al 1992) comprise
varved clay deposits to 13.5 m to 15.3 m depth overlying dense sandy and silty moraine.
The varved clay profile comprises a 2 m thick fissured crust overlying soft, high
plasticity clays to 10 m depth. Below 10 m depth the clay is of firm to stiff strength
consistency (undrained shear strength of 30 to 50 kPa) and the varves are thicker,
sandier and siltier. Below the heavily over-consolidated crust the over-consolidation
ratio (OCR) decreases from about 12 at 2 m depth to 4 at 4 m depth to less than 2.4
below 8 m depth.
The constructed embankment is of 6 m maximum height with very flat slopes
(Figure C2.11). It was constructed in approximately 40 days (Day - 40 to Day 0 in
Figure C2.12).
The excess pore water pressure response with time of the piezometers under the
centreline, upper slope and mid berm are shown in Figure C2.12.

Figure C2.11: Section A of the Olga-C test embankment with no wick drains (Lavallée
et al 1992).
Appendix C Page C19

a) 140 140
A105 (centre, 10.5 m)
A106 (centre, 7.5 m)
120 A107 (centre, 4.5 m) 120

Applied Embankment Load (kPa)


Excess Pore Water Pressure
A108 (centre, 2 m)
Applied Load (kPa)
100 100

80 80
(kPa)

60 60

40 40

20 20

0 0
-80 -60 -40 -20 0 20 40 60 80 100 120
Time (Days)

b) 140
A102 (slope, 7.5 m)
140

A103 (slope, 4.5 m)


Excess Pore Water Pressure (kPa)

120 120

Applied Embankment Load (kPa)


A104 (slope, 2 m)
Applied Load (kPa)
100 100

80 80

60 60

40 40

20 20

0 0
-80 -60 -40 -20 0 20 40 60 80 100 120
Time (Days)

c) 140 140
Excess Pore Water Pressure (kPa)

120 120
Applied Embankment Load (kPa)

100 C120 (berm, 7.5 m) 100

C119 (berm, 10.5 m)


80 80
C121 (berm, 4.5 m)
Applied Load (kPa)
60 60

40 40

20 20

0 0
-80 -60 -40 -20 0 20 40 60 80 100 120
Time (Days)
Figure C2.12: Olga-C, excess pore water pressure response with time at Section A (a)
below the centre line, (b) below the upper slope and (c) below the berm.
Appendix C Page C20

-3

-1.5
Fill Height = 3 m (Day -12)
0
Fill Height = 6 m (Day 0)

1.5 Fill Height = 6 m (Day 7)

Fill Height = 6 m (Day 26)


3
Fill Height = 6 m (Day 110)
4.5
Depth (m)

7.5

10.5

12

13.5 Note: 0 m depth is the


original ground surface
15
-3 -2 -1 0 1 2 3 4
Vertical Inclination Angle (rad x 102)

Figure C2.13: Olga-C, vertical inclination angle profiles at various times at Inclinometer
INC1, located at about mid berm, at Section A.

Figure C2.13 presents the plot of vertical inclination angle for Inclinometer INC1
positioned at about mid berm (Figure C2.11). It shows the development of what is
considerer to be yielding between 5.5 m and 11.5 m depth and possibly the development
of localised failure between 6 m and 8.5 m depth.
In the context of this observation, the excess pore water pressure response of the
piezometers after the end of construction (at Day 0) is considered to represent:
• Rapid dissipation of excess pore water pressure immediately below the crust
(Piezometers A104 and A108) and at 4.5 m depth under the berm (Piezometer
C121) of the over-consolidated clay foundation. The dissipation of excess pore
water pressures in these piezometers starts shortly after the end of construction at
Day 0.
• Piezometers A105 and C119, located in the firm to stiff clays at 10.5 m depth, show
a similar response except that the excess pore water pressure increases for about 10
days before decreasing. It is considered that the foundation at this depth is still in an
over-consolidated condition.
Appendix C Page C21

• The remaining piezometers at depths of 4.5 m and 7.5 m depth under the crest and
slope (Piezometers A106, A107, A102 and A103) show a slow increase in excess
pore water pressure for 50 to 70 days before decreasing. It is considered that this
response represents yielding of the normally consolidated soft clays. The increase
in excess pore water pressure is considered to be due to the reduction of the limit
state surface (refer Section 4.4.2 of Chapter 4).

Leroueil (1996) report that the vertical strains 300 days after construction were
relatively uniform throughout the varved clay, with an average of 6% vertical strain at
Section A (5.8% at 7.7 m depth and 8.5% at 10.4 m depth). Leroueil concluded that the
post construction settlement behaviour and pore water pressure response at Section A of
the Olga-C embankment indicates that the undrained creep and consolidation processes
occur simultaneously. In Piezometer A106 (7.5 m depth below the centre) the excess
pore water pressure at 300 days after construction was still greater than the pressure on
Day 0 at end of construction, yet almost 6% vertical strain has occurred in this region of
the foundation.

2.3 QUEENSBOROUGH TEST EMBANKMENT


The Queensborough test embankment (Figure C2.14) is located in North Kent, UK and
was constructed in 1976 over soft alluvial soils of the Thames estuary.
The subsurface conditions at the Queensborough test site (Jardine and Hight 1987)
consist of a firm, laminated silty clay crust of 0.5 to 2.3 m thickness overlying soft to
very soft alluvial silty clays to 8.5 to 10.3 m depth. The soft clays are described as
being relatively homogeneous. The OCR ranges from approximately 4 in the crust, to
2.5 below the crust decreasing to 1.1 below 4.5 to 6 m depth. London Clay underlies
the soft alluvial clays.
The embankment comprised an initial layer of sand and gravel to act as a drainage
medium overlain by chalk filling. The rate of construction was approximately 0.07
m/day. Construction was halted on Day 108 following observation of cracking near the
centre of the embankment. The embankment height at this time was a maximum of 2.9
m. Due to concerns over potential failure of the embankment, the fill height was
reduced some 150 mm and a toe surcharge was placed in critical areas from about Day
114.
Appendix C Page C22

Figure C2.14: Queensborough test embankment, section of test embankment showing


instrumentation (Jardine and Hight 1987).

Figure C2.15 presents the excess pore water pressure response (during and post
construction) versus time of two piezometers located in lower, very lightly over-
consolidated region (OCR approximately 1.1) of the soft clay alluvium. Piezometer
P2/1 is located under the toe of the embankment and Piezometer P5/2 is located under
the centre of the embankment. Pore water pressures in both these piezometers were
observed to rise for the six days following halting of construction (Day 108 to 114).
Pore water pressures were observed to decrease after about Day 124.
Figure C2.16 presents the observations of settlement under the centre of the
embankment and maximum lateral deflection at the toe of the embankment (from
inclinometer I8) during and post construction. It is evident that large lateral
displacements at the embankment toe occurred during the later stages of embankment
construction and continued after the end of construction. It is considered that these
displacements are a result of plastic deformation within a localised failure zone;
however, as no records of the inclinometer were published for I8 it was not possible to
determine the location of the localised failure zone. A reduction in the rate of lateral
displacement was observed after about Day 125, coincident with the time at which the
excess pore water pressures started to decrease in Piezometers P2/1 and P5/2 (Figure
C2.15).
Jardine and Hight (1987) commented that the ratio of ∆y m / ∆ s = 0.65 for the first

year post construction reducing to about 0.2 thereafter. They considered it to be


significantly affected by undrained creep deformations. From Figure C2.16 it is evident
that the ratio ∆y m / ∆s post construction is closer to 1 for about the first 40 days after
Appendix C Page C23

construction, and by about 90 days after construction (Day 200) had decreased to
approximately 0.25.

a) 60

P2/1 (toe, RL -5.6 (8 m depth))


Excess Pore Water Pressure (kPa)

P5/2 (centre, RL -6.2 (8 m depth))


50
Undrained Response 114
124
170
617 108
40 106

100
101
30
97 114
124
170
91
108
106 617
20
100
85 101
78 80
97
68 91
10
85
78

0
0 10 20 30 40 50 60
Applied Load from Fill (kPa)

b) 60
Excess Pore Water Pressure / Applied

50

40
Load (kPa)

30

20
P2/1 (toe, RL -5.6 (8 m depth))
P5/2 (centre, RL -6.2 (8 m depth))
10
Applied Load (kPa)

0
0 100 200 300 400 500 600 700
Time (Days)
Figure C2.15: Queensborough test embankme nt, excess pore water pressure response
versus (a) applied embankment load and (b) time.
Appendix C Page C24

Settlement, S6/5 (RL 0.6 m (0.7 m depth))


20
Settlement and Displacement (mm)
Settlement, S6/4 (RL -1.2 m (2.5 m depth))
40 Lateral Displacement ym, I8 (toe)

60

80
End of
construction
100

120

140

160
0 100 200 300 400 500 600 700
Time (Days)
Figure C2.16: Settlement at the centre and maximum lateral displacement at the toe of
the Queensborough test embankment.

Several other important observation made by Jardine and Hight (1987) were:
• A yield condition was reached at all depths in the soft clay alluvium below the
desiccated crust, with the deeper layers first to yield. From the available
information, the threshold height at Piezometer P5/2 was estimated at 0.7 m
(equivalent to 14 kPa applied load).
• Settlements during construction were small and were generally concentrated within
the upper layers of the soft alluvium.
• The post construction settlements in the lower strata far exceeded those accumulated
by pore water pressure dissipation, indicating the significance of plastic deformation
in the localised failure zones and undrained creep in the initial post construction
period.
Appendix C Page C25

3.0 CASE STUDY EXCEPTIONS TO GOOD INDICATORS


OF AN IMPENDING FAILURE
This section discusses further those case studies of embankments on soft ground
identified in Section 4.3 of Chapter 4 for which:
• The excess pore water pressure response was an exception to the normal trend of
most of the failure case studies analysed (refer Table 4.3 of Section 4.3.2).
• Deformation was a poor indicator of an impending failure condition or the change in
deformation behaviour had alternative explanations (refer Table 4.4 of Section
4.3.3.6).

3.1 EXCESS P ORE WATER P RESSURE R ESPONSE

3.1.1 Field Estimation of the Pre-Consolidation Pressure

Piezometer A3 at James Bay (Dounias and Tournier 1975) was identified as an


exception with respect to the large difference between the effective critical stress, σ 1′, crit ,

estimated from the field pore water pressure response and the pre-consolidation pressure
estimated by laboratory oedometer testing, σ ′p (Table 4.2 of Section 4.3.2 of Chapter

4). The effective critical stress was significantly lower than σ ′p .

A similar pattern of behaviour between the effective critical stress and pre-
consolidation pressure was observed for Piezometers A1 and A2 (at shallower depth) at
this site. At greater depth, Piezometer A4 (10.7 m depth), the effective critical stress
was within 1% of the pre-consolidation pressure.
Leroueil et al (1978a) considered that the observation was related to the effective
stress path for heavily over-consolidation soils (OCR > 2.5), which intersects the limit
state surface at less than the pre-consolidation pressure (Figure C3.1); i.e. at Point P1.
Therefore, for heavily over-consolidated soils the effective critical stress, σ 1′, crit ,

estimated from the field pore water pressure response would under-estimate the pre-
consolidation pressure, PC, as shown in Figure C3.1.
Appendix C Page C26

Figure C3.1: Interpretation of effective stress path for heavily over-consolidated clays
under embankments (Leroueil et al 1978a)

3.1.2 Excess Pore Water Pressure Response for Normally Consolidated Conditions

For the piezometric response during normally consolidated conditions ( B2 ), the


analysed case studies indicate that in most cases the excess pore water pressure
observations maintain a B 2 ≈ 1 up to the failure height (Table 4.2 in Section 4.3.2 of
Chapter 4). Four case studies exhibit excess pore water pressure observations other than
B 2 ≈ 1 ; Kalix, Mastemyr, Muar-EC and Cubzac-A.
For Mastemyr (foundation of very soft, quick, leached marine clay deposit) the
initial excess pore water pressure response (Figure 4.21a in Section 4.3.2 of Chapter 4)
indicates a B1 value of 1.0 reducing to 0.5 up to about 17 kPa applied load. At greater

than about 28 kPa applied load, the ∆B response in P1 is then significantly greater than
1. This response is similar for piezometers down to 6 m depth at this site.
Leroueil et al (1978a) consider that the initial effective stress path to be similar to
path OYEc in Figure C3.2a, where it intersects with the limit state surface above the
critical state line. They comment that of the 30 case studies they analysed this was the
only case for which this behaviour was observed, indicating its unusual occurrence.
Leroueil et al could not establish the reasons for this.
The reason that the effective stress path follows this path may possibly be due to the
very low permeability of the foundation.
The consistent pore water pressure response of ∆B > 1 for all piezometers under the
embankment centreline in the upper 6 m depth at applied embankment loads greater
Appendix C Page C27

than 28 kPa may be indicative of undrained strain weakening due to partial breakdown
of the soil structure throughout at least the upper 6 m of the very soft clay deposit.

Ec = end of construction
Figure C3.2: Effective stress paths and pore water pressure generation associated with
different types of yield conditions (Leroueil and Tavenas 1986).

The piezometers under the embankment centreline at Kalix (foundation of soft to very
soft, moderately sensitive, highly plastic, marine deposit of silt/organic clay) showed a
similar response to that at Mastemyr for a depth range from the below the crust to a
depth of at least 7.5 m. This response may also indicate strain weakening due to partial
breakdown of the soil structure over this depth range.
In the case of Mastemyr and Kalix it is considered that the effective stress path is
similar to that shown in Figure C3.2a (path OYEc) where the effective stress path
intersects the limit state surface above the critical state line. On reaching the limit state
surface (Point Y) a localised failure condition is reached and with further increase in
applied load, strain weakening occurs resulting in the observation of excess pore water
pressures exceeding the increase in total vertical stress (i.e. ∆B greater than 1).
For both Mastemyr and Kalix the excess pore water pressure response equivalent to
∆B greater than 1 was observed in a number of piezometer and is considered to be
indicative of a partial breakdown over a significant portion of the foundation. What is
not clear is why a general failure did not occur once a significant portion of the
Appendix C Page C28

foundation had reached a failure condition. In the case of the Kalix embankment the
foundation comprised a highly plastic sensitive clay, and what influence this may have
had is uncertain
For the Muar-EC embankment, the excess pore water pressure response of
Piezometer P4 (Figure 4.21b in Section 4.3.2 of Chapter 4) was the only piezometer for
this case study having a B2 > 1 . The response of other piezometers under the centre of
the embankment, once normally consolidated conditions have been reached, maintain a
∆B value of close to 1. It is uncertain what has caused this observed behaviour at
Piezometer P4.
For the Cubzac-A embankment, the observed ∆B response for Piezometer P112
located at 4 m depth under the embankment centreline (Figure 4.21a in Section 4.3.2 of
Chapter 4), is greater than 1 above an applied embankment load of 52 kPa. A similar
response was measured in the underlying piezometer at 6 m depth. It is uncertain what
has caused this observation. It is possible that the response is due to time effects
(discussed in Section 4.4.2 of Chapter 4) or else is due to general strain weakening
throughout the depth of the soft to very soft, highly plastic clay deposit.

3.2 P RE-F AILURE D EFORMATION RESPONSE

3.2.1 Rio Case Study

For the Rio test embankment (foundation of high plasticity clays of low sensitivity), the
vertical displacement at and beyond the toe of the embankment and lateral surface
displacement at the embankment toe are considered as reasonable to good indicators of
an impending failure condition. However, the abrupt changes in deformation behaviour
that occur at a relative embankment height of 89% of the failure height are coincident
with the observation of cracking in the embankment. Therefore, the changes in
deformation behaviour are not explainable only by being close to the failure height as
the cracking of the embankment is considered to have had an effect on the increased
rates of displacement observed due to changes in the effective stress conditions in the
foundation. It could also be said that the lateral displacement probably caused the
cracking.
The lateral displacement (at depth) at the embankment toe, from the inclinometer
observations, was considered as a poor indicator of an impending failure for the Rio test
Appendix C Page C29

embankment. The plots of vertical inclination angle from the inclinometer did not
indicate development of the localised failure zone until the embankment had cracked at
an embankment height of 89% of the failure height.
It is not certain why the inclinometer did not indicate the development of the
localised failure zone prior to failure. A possible reason is that it was not until the
embankment cracked at 2.5 m height that the stresses within the foundation were such
that a localised failure condition was reached under the shoulder of the embankment and
large plastic deformations occurred. In addition, the soft clay may be typical of a
classical ductile clay that is not strain weakening on shearing, however, greater straining
in the more highly stressed regions under the embankment slope should still have been
evident. Another possible influence is considered to be the influence of the two small
embankments constructed at the toe main embankment on the stress conditions within
the foundation. The purpose of these small embankments, constructed only 20 m apart,
was to contain the failure within the monitored portion of the site.

3.2.2 Kalix Case Study

For the Kalix embankment (foundation of highly plastic, sensitive clays and silts) all
indicators provided a reasonable to good indication of the impending failure condition
(the vertical deformation at and beyond the embankment toe, and the lateral surface and
at depth displacement at the embankment toe). However, the deformation behaviour is
not explainable only by being close to the failure height.
A significant change in the deformation behaviour occurred at embankment heights
of 70 to 90% of the failure height, and this was coincident with the threshold fill height
(heave movements at the embankment toe, large increase in the rate of lateral
displacement at the embankment toe, both at the surface and at depth were measured).
Whilst the abrupt change is considered a good indicator, it is also explainable by
deformation changes associated with reaching the threshold fill height.
As discussed in Section 3.1.2 (of Appendix C), it was considered that the effective
stress path for Kalix might have intersected the limit state surface above the critical state
line resulting in the abrupt change in ∆B from 0.4 to 1.7. As a consequence, strain
weakening due to partial breakdown of the soil structure may have occurred over a
significant depth of the upper soft to very soft clay. The change in deformation
Appendix C Page C30

behaviour at the embankment toe is generally coincident with this change in pore water
pressure response.

3.2.3 King’s Lynn Case Study

At the King’s Lynn embankment (foundation of soft to very soft, high plasticity marine
deposit of low sensitivity), an impending failure condition is not clearly identifiable
from the plot of lateral surface displacement at the embankment toe versus relative
embankment height (Figure 4.25 in Section 4.3.3.3 of Chapter 4). However, the
observations of lateral surface displacement at the toe and cracking of the embankment
did provide an indication of the impending failure.
On the morning of the day prior to failure, fine longitudinal cracks were observed on
the crest of the embankment (at a relative height of 92% of the eventual failure height)
and 10 mm of lateral displacement was recorded overnight. An additional fill lift of
0.55 m thickness was then placed during which time 14 mm of lateral surface
displacement was recorded at the toe. In the next 6 to 8 hours of that day a further 45
mm (or 55% of the total lateral surface displacement recorded to completion of fill
placement) was recorded, providing a clear indication of the impending failure. Failure
occurred at about 8 am the next morning.
The plot of incremental vertical inclination angle (from the inclinometer at the toe of
the embankment), Figure 4.26b in Section 4.3.3.4 of Chapter 4, clearly shows the
development of the localised failure zone once the embankment had been constructed to
its eventual failure height of 6.7 m.

4.0 POST-FAILURE DEFORMATION ANALYSIS


The Khalili et al (1996) method for estimating the post-failure deformation of an intact
slide requires the surface of rupture to be known and knowledge of the strength
properties at and post failure. For each of the case studies analysed, limit equilibrium
analysis was undertaken to calculate a factor of safety of 1 for the initial failure and for
the modelled surface of rupture to be a good representation of the actual (refer Section
4.3.4 of Chapter 4 for further details on the modelling). The limit equilibrium analysis
was then re-run using the modelled surface of rupture, but with residual strength
parameters, to calculate the residual factor of safety, FSresidual.
Appendix C Page C31

In all four cases, use of the remoulded vane strengths resulted in very low residual
factors of safety and post-failure movement predictions significantly greater than the
observed movement. On evaluation, it was evident that the remoulded vane strengths
indicated that the amount of strain weakening to residual strength was much greater
than that measured in undrained triaxial tests. Therefore, estimates of the residual
strength using the vane were considered to under estimate the actual values.
In the case of Rio and Muar-F the large strain undrained shear strength was also
obtained from an estimate of the drained residual friction angle. The post failure
deformation predictions based on the estimates of the residual undrained strength profile
from laboratory test results (St. Alban-A) and from the residual friction angle (Rio and
Muar-F) were much closer to the observed post-failure deformation behaviour.
Further discussion of each case study analysed is presented in the following sub-
sections.

4.1 RIO TEST EMBANKMENT, RIO DE JANIERO


The Rio test embankment was constructed over a high plasticity marine clay (Figure
C4.1). Field vane tests indicated the remoulded strength to range from 1.5 to 5 kPa and
the clay to be of low sensitivity ( S I ≈ 3 , where SI = sensitivity). Laboratory undrained
triaxial testing indicate the soft clay foundation to be highly ductile with deviatoric
stress still increasing at axial strains of 8 to 10%.
A weathered, fissured crust of 2.4 m thickness overlies the soft clay foundation.
The embankment was constructed of loosely placed silty sand.
During construction, cracking of the embankment was initially observed at an
embankment height of 2.5 m. After raising to 2.8 m the next day, severe cracking was
observed but no slip failure (2.8 m was deemed to be the failure height). After several
days still no slip failure was observed so the embankment was raised to 3.1 m height,
which resulted in additional cracking but still no slip failure. The next day, during
placement of the next lift on the embankment, a general slip failure “slowly” occurred.
The total crest settlement and heave beyond the embankment toe of the slip was
approximately 680 mm.
Adopting the remoulded strength profile from the vane tests and modelling the
cracked embankment resulted in a residual factor of safety of 0.43 for the reported
surface of rupture (Figure C4.1). This gave a crest settlement of 1.1 m for the slow
Appendix C Page C32

model and 2.1 m for the rapid model, both of which are well above the actual crest
settlement of 680 mm. It is considered that this is due to the under prediction of the
residual strength from the remoulded vane test and is typical of all the embankment
cases analysed.

Figure C4.1: Pre and post-failure geometry of Rio test embankment (Ramalho-Ortigao
et al 1983b)

The undrained triaxial tests on normally consolidated samples of the soft clay from the
Rio site (Almeida 1982) indicate strain-hardening behaviour up to axial strains of 8 to
10%. Therefore, strength reductions to 30% of the peak strength (as indicated from the
remoulded vane tests) would appear to be a significant under estimate of the residual
strength. A better estimate of the residual strength (Sur) from the laboratory results was
considered to be 80% of the peak strength. This was estimated by adopting a residual
friction angle (φ r′ ) of 12 degrees, based on the clay fines content and plasticity of the
soft clay (Fell et al 1992), and estimating the undrained strength profile from Equation
C4.1 (Equation 4.8 in Section 4.3.5.2 of Chapter 4).

S ur = σ ′p tan φr′ (C4.1)


Appendix C Page C33

where σ ′p is the pre-consolidation pressure of the soft clay, φ r′ the drained residual

friction angle and Sur the residual undrained strength.


The residual factor of safety adopting the residual strength profile calculated from
Equation C4.1 was 0.70. This resulted in a crest settlement of 590 mm for the slow
model and 1140 mm for the rapid model, much closer to the observed crest settlement
of 680 mm.

4.2 M UAR-F TEST EMBANKMENT, M ALAYSIA


The Muar-F test embankment was constructed over a soft, high plasticity marine clay
foundation (Figure C4.2). A weathered crust of 2 m thickness overlies the soft clay
foundation and the embankment was constructed of compacted, highly plastic
decomposed granite.

Figure C4.2: Pre and post-failure geometry of Muar-F test embankment (Brand and
Premchitt 1989)

Field vane tests indicated the remoulded strength of the soft foundation to range from
2.5 to 7 kPa and the clay to be of low sensitivity ( S I ≈ 4 where SI = sensitivity).
Effective stress path plots of laboratory undrained triaxial testing (Malaysian Highway
Authority 1989a) indicate the soft clay foundation (at close to the in-situ stress
conditions) to show little strain weakening and a large strain deviatoric stress of about
90% of the peak deviatoric stress.
During construction, cracking of the embankment was initially observed at an
embankment height of 5.4 m with increases in the crack width occurring later that
afternoon and over-night. The next day the embankment was raised to 5.6 m height and
failure occurred shortly after placement of this layer. The failure occurred rapidly
Appendix C Page C34

(within approximately 3 minutes) and the total crest settlement was approximately 1.4
m.
Adopting the remoulded strength profile from the vane tests and modelling the
cracked embankment resulted in a residual factor of safety of 0.40 for the reported
surface of rupture (Figure C4.2). This resulted in a crest settlement of 0.78 m for the
slow model and 1.53 m for the rapid model. The observed crest settlement of 1.4 m is
close to the prediction using the rapid model.
However, it is suspected that the reasonable estimate of the rapid model to the
observed movement is somewhat fortuitous, as the remoulded strength from the vane
would indicate the remoulded strength to be 25% of the peak strength. The results of
the laboratory undrained triaxial strength tests do not indicate the remoulded strength to
be as low as this.
Adopting a residual friction angle of 12 degrees for the soft clay foundation, based
on its clay fraction and plasticity, the residual undrained strengths are estimated to be in
the range 7.5 to 13.5 kPa (from Equation C4.1), and the residual factor of safety
estimated at 0.71 from limit equilibrium analysis. This resulted in a crest settlement of
0.38 m for the slow model and 0.75 m for the rapid model, significantly lower than the
actual movement.
It is suspected that given the large radius (26 m) and relatively shallow depth (8.2 m
maximum) of the surface of rupture (Figure C4.2) that the actual surface of rupture may
have had a significant translational component. This would result in some internal
brittleness of the slide that would explain the “very loud noise” at failure and relatively
rapid movement of a soft clay foundation of low sensitivity. In addition, a wedge type
settlement of the crest would have occurred for the compound failure surface geometry
with rotational back-scarp, and possibly greater crest movements than could be
predicted by the Khalili et al (1996) model.

4.3 ST. ALBAN-A TEST EMBANKMENT, QUEBEC, CANADA


The St. Alban-A test embankment was constructed over a soft to very soft, low
plasticity leached marine clay/silt (Figure C4.3). A weathered, fissured crust of 1.5 to
1.8 m thickness overlies the soft clay foundation and the embankment was constructed
of compacted, uniform sand.
Appendix C Page C35

Figure C4.3: Pre and post-failure geometry of St. Alban-A test embankment (La
Rochelle et al 1974)

Laboratory triaxial testing indicates the soft foundation to be cemented, sensitive and
brittle (La Rochelle et al 1974). The undrained triaxial tests of the soft clay foundation
(at close to the in-situ stress conditions) show significant strain weakening with a large
strain deviatoric stress of about 75% of the peak deviatoric stress. Field vane tests
indicate the remoulded strength to range from 1 to 3 kPa (significantly lower than that
from the laboratory undrained triaxial tests) classifying the soft foundation as highly
sensitive (SI of 8 to 22) when compared to the peak vane results.
During construction, no cracking of the embankment was observed prior to failure.
The failure occurred rapidly (within approximately 1 minute) at an embankment height
of 4 m. The total post failure crest settlement was approximately 1.15 m and toe heave
1.2 m.
Adopting the remoulded strength profile from the vane tests a residual factor of
safety of 0.26 was obtained by limit equilibrium analysis for the reported surface of
rupture (Figure C4.3). This resulted in a crest settlement of 2.3 m for the slow model
and 4.6 m for the rapid model, significantly greater than the observed crest settlement of
1.15 m.
Adopting a residual strength of 75% of the peak strength (equivalent to the reported
large strain undrained shear strength from the laboratory undrained triaxial tests) a
residual factor of safety of 0.75 is obtained from limit equilibrium analysis. This
resulted in a crest settlement of 0.79 m for the slow model and 1.53 m for the rapid
model, significantly closer to the observed crest settlement of 1.15 m. It would have
been expected that the rapid model would have more closely predicted the observed
movement than the results indicate given the rapid post failure velocity of the failed
Appendix C Page C36

slide mass. It is considered that the sensitivity of the residual factor of safety to the
strength parameters adopted is probably why the prediction was not closer.

4.4 P ORTSMOUTH TEST EMBANKMENT, N EW HAMPSHIRE, USA


The Portsmouth test embankment was constructed over a soft, medium plasticity
leached marine silty clay (Figure C4.4). A weathered, fissured crust of 2.4 m thickness
overlay the soft clay foundation and the embankment was constructed of clean sand.
The remoulded vane test results indicate the soft foundation to be sensitive ( S I ≈ 10 )
with remoulded strengths ranging from 1.4 to 7 kPa. Laboratory triaxial undrained
testing, as discussed by Ladd (1972), indicates the foundation has significant strength
anisotropy. It is suspected that it would also exhibit brittle, and possibly partially
cemented, undrained shear strength behaviour with significant post peak strain
weakening, although, no stress-strain plots are published by Ladd.
During construction no cracking of the embankment was observed prior to failure at
an embankment height of 6.55 m. The failure is suspected to have occurred rapidly.
The total crest settlement was approximately 3.3 to 3.4 m with a similar amount of
heave movement beyond the embankment toe.

Figure C4.4: Pre and post-failure geometry of Portsmouth test embankment (Ladd 1972)

Adopting the remoulded strength profile from the vane tests a residual factor of safety
of 0.27 was obtained by limit equilibrium analysis for the reported surface of rupture
Appendix C Page C37

(Figure C4.4). This resulted in a crest settlement of 3.4 m for the slow model and 6.4 m
for the rapid model. Given the suspected rapid post-failure movement, the predicted
rapid movement is significantly greater than the observed crest settlement of 3.3 to 3.4
m. Once again it is suspected that the remoulded vane strength significantly under
estimates the residual strength. However, no laboratory triaxial undrained test data in
the form of stress-strain plots were published by Ladd (1972) to allow assessment of the
large strain undrained shear strength.
APPENDIX D

Case Studies of Slope Failures


in Embankment Dams and in
Cut Slopes in Heavily Over-
Consolidated High Plasticity
Clays
Appendix D Page i

TABLE OF CONTENTS

1.0 CASE STUDY INFORMATION ON LANDSLIDES IN


EMBANKMENT DAMS............................................................. D1

2.0 FAILURES IN CUT SLOPES IN HEAVILY OVER-


CONSOLIDATED HIGH PLASTICITY CLAYS .................. D22
2.1 Cut Slopes in London Clay .......................................................................... D22
2.2 Cut Slopes in Weathered Upper Lias Clay................................................... D24
2.3 Terminology Used in Tables ........................................................................ D26

LIST OF TABLES
Table D1.1: Case studies of slope instability of embankment dams – slides during
construction. ............................................................................................................ D3
Table D1.2: Case studies of slope instability of embankment dams – slides in the
upstream slope during drawdown. ........................................................................ D13
Table D1.3: Case studies of slope instability of embankment dams – post-
construction slides in the downstream slope......................................................... D18
Table D2.1: Summary of material properties of London clay*1 .................................. D23
Table D2.2: Summary of material properties of weathered Upper Lias clay*1 ........... D26
Table D2.3: Case studies of failures in cut slopes in London clay .............................. D29
Table D2.4: Case studies of failures in cut slopes in weathered Upper Lias clay ....... D31
Appendix D Page D1

1.0 CASE STUDY INFORMATION ON LANDSLIDES IN


EMBANKMENT DAMS
The case study information for embankment dams is presented in the following tables:
• Table D1.1 for slides in embankment dams during construction;
• Table D1.2 for post-construction slides in the upstream slope of embankment dams
during drawdown; and
• Table D1.3 for slides in the downstream slope of embankment dams that occur post-
construction.

An explanation of some of the terminology used in these tables is as follows:


• Some general terms or symbols used and their meaning:
− (?) = unsure of detail, but thought to be likely
− unk = unknown
− “-” = not required or not appropriate
− H to V refers to horizontal to vertical slope angle; i.e. 2H to 1V.
− Su = undrained shear strength
− Spec. = specification.
− SMP = surface monitoring point.
• Under the section embankment zoning and classification;
− “Zoning” refers to the embankment zoning classification (refer Figure 1.6 in
Section 1.3.3 of Chapter 1)
− “Classification” refers to the embankment and filter classification (refer
Figure 1.5 in Section 1.3.3 of Chapter 1)
− “Core Width” refers to the thickness classification of the core
− “Core Type” refers to the classification of the core in accordance with the
Australian soil classification system (Australian Standard AS 1726-1993
“Geotechnical Site Investigations”).
• Slide location refers to the location of the surface of rupture, being either:
− Emb = slide through the embankment only
− Emb & Fndn = slide through the embankment and foundation
• For the material properties:
Appendix D Page D2

− ASCS Classn = Australian soil classification system (Australian Standard AS


1726-1993).
− % fines = percentage finer than 75 micron (i.e. percent silt and clay fraction)
− CF (%) = percentage of clay fines (i.e. less than 2 micron)
− PI (%) = plasticity index
− “Density Ratio” = ratio of the field dry density to Standard maximum dry
density (SMDD), in accordance with Australian Standard AS1289.5.1.1 or the
Standard Proctor test. Unit weights or densities are quoted where given.
− Moisture Content = moisture content in relation to Standard optimum
moisture content (OMC); dry being dry of OMC, wet being wet of OMC.
• Slide classification refers to the classification of the shape of the surface of rupture
(e.g. translational, rotational or compound) using the landslide classification system
of Hutchinson (1988).
• For the post-failure deformation behaviour:
− Velocity category refers to the IUGS (1995) velocity categories.
− The degree of break-up is a somewhat subjective assessment of the post-
failure break up of the slide mass.
Appendix D Page D3

Table D1.1: Case studies of slope instability of embankment dams – slides during construction.
Embankment Zoning and CONSTRUCTION
Classification Construction Embankment Dimensions
Date of Slide
Name Location Geology References
Failure Location
n Core Core Year Rate Height, He Length, Ratio Upstream Downstream
Type Class Comments
Width Type Started (mm/day) (m) Le (m) Le /H e Slope Slope

Lateral bulging during construction.


Linell & Shea (1960)
Otter Brook USA, New Metamorphic, Homogeneous earthfill dam with
Aug-57 Homog 1,1,1 c-vb SC Emb 1957 470 (max) 40.5 unk - 2.5H to 1V 2.5H to 1V Sherard et al (1963)
Dam England glaciation downstream chimney drain and foundation
Walker and Duncan (1984)
filter

Lateral bulging and cracking observed


CL/CH 360 (max), 2.5 to 3.5H
Truscott Dam USA, Texas 03-Dec-81 Homog 1,1,1 c-vb Emb unk 1980 30 4720 157.3 3H to 1V during latter stages of construction within Walker and Duncan (1984)
(?) last 6.2 m fill. to 1V
closure section (6.2 m below design level).

Lateral bulging and cracking observed


USA, CL/CH 230 (closure 2.5 to 4H to
Skiatook Dam Aug-82 Homog 1,1,1 c-vb Emb unk 1977 33.8 1090 32.2 3 to 4H to 1V during construction within closure section Walker and Duncan (1984)
Oklahoma (?) section) 1V
(when 11 m below design level).

Quartz diorite - Lateral bulging of downstream slope when Villegas (1984)


2.5 to 3.5H to
Troneras Dam Columbia 12-Apr-62 Homog 1,1,1 c-vb ML Emb deeply weathered 1960 360 to 1330 31.5 366 11.6 3 to 4H to 1V construction rate in excess of 1 m/day Li (1963)
1V
profile. (upper 8 m of embankment). Li (1967)

45 (35 Lateral bulging (incipient failure) of


Punchina when downstream slope. Occurred when
Columbia 3-Apr-80 Homog 2,0,0 c-vb SM Emb Quartz diorite 1980 570 to 1500 350 7.8 3.4H to 1V 2H to 1V Villegas (1982)
Cofferdam movement constructing downstream section at very
observed) rapid rate(approx 1.5 m/day).

Penman (1986)
Reconstruction of gorge section in 1884
Waghad India Apr 1884 Homog 0,0,0 c-vb CH/MH Emb unk 1881 240 (g) 32 1270 39.7 3H to 1V 2H to 1V Nagarkar et al (1978)
following overtop in 1883.
Nagarkar et al (1981)

Sedimentary - 2.5 to 4H to Failure within closure section on upstream Humphrey & Leonard (1986,
Lake 100 (closure 3 to 4H to 1V
USA, Illanois 17-Sep-70 Homog 1,1,1 c-vb CL Emb shale, sandstone, 1966 33 1034 31.3 1V (with slope. Observed when construction 1988)
Shelbyville section) (with berm)
coal, limestone berm) complete. Crooks & Becker (1988)

Scout Sedimentary - 30 (but


Instability of upstream slope. Only
Reservation USA, Ohio Sep-84 Homog 1,0,1 c-vb CL Emb shale, claystone 1983 suspect much 13.7 245 17.9 2H to 1V 3H to 1V Mann and Snow (1992)
observed when construction completed.
Dam and sandstone quicker)

From May 1941 construction rate


Puddle 25 increasing Banks (1948)
Volcanic, overlain 21.3 (at increased to approx 85 mm/day (4 fold
Muirhead UK, Scotland Sep-41 core 8,0,1 - CL/CH Emb 1940 to 85 (85 at 631 29.6 2.5H to 1V 3H to 1V Skempton (1990)
by Boulder Clay failure) increase in volume). 5 m below design
earthfill failure) Penman (1986)
height at failure.

Zoned
Most of construction in 1935, completed
Dam FC10 Jan-37 Earth and 4,0,0 c-vb CH Emb unk 1935 60 (g) 19.5 290 14.9 2 to 4H to 1V 3 to 4H to 1V
Dec 1936. Failure shortly after completion.
rockfill

Zoned
34.8 (at Failed during construction when 5.2 m
Acu Dam Brazil 15-Dec-81 Earth and 4,1,1 c-tk CH (?) Emb unk 1980 (?) unk unk - 2.5H to 1V 2.5H to 1V Penman (1985)
failure) below design crest level.
rockfill
Appendix D Page D4

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 2 of 10)
EMBANKMENT MATERIALS
Material 1 (Core) Material 2 (Shoulder fill)
Name
ASCS Density Moisture ASCS Density Moisture
Source n % fines CF (%) PI (%) Compaction Comments Source n % fines CF (%) PI (%) Compaction Comments
Class Ratio Content Class Ratio Content

0.6% dry Rolled, well compacted. 200


Otter Brook SC - clayey Large variation in moisture content at
Glacial till 45 12 15 100.5% (20% > 1% mm layers, 6 passes of - - - - - - - - -
Dam fine sand placement.
wet) sheepsfoot roller (16 tonne)

Investigation indicates fill of stiff to


Variable,
Sedimentary, Suspect rolled, well very stiff strength consistency (Su ~
Truscott Dam CH (?) unk unk unk unk some wet - - - - - - - - -
clay shales compacted 55 to 190 kPa). Stiff zones
zones
associated with wetter zones.

Investigation indicates fill of stiff to


Suspect rolled, well very stiff strength consistency. Lower
Skiatook Dam unk CL/CH (?) unk unk unk unk 16 to 20% - - - - - - - - -
compacted strength zones (Su ~ 65 to 70 kPa)
associated with wetter zones in fill.

14.4
4% wet Very rapid rate of construction.
Residual and kN/cu.m. Rolled, well compacted. Pad
Troneras Dam ML 60 to 80 5 to 15 8 to 12 (avg), up to Placed well wet OMC, low undrained - - - - - - - - -
saprolitic soils (dry foot rollers
10% wet shear strength.
density)

Very rapid rate of construction.


Punchina Residual and 4.1% wet Rolled, well compacted. Pad Placed well wet OMC. Material had
SM 43 6 10.2 95.4% - - - - - - - - -
Cofferdam saprolitic soils (avg) foot rollers low undrained shear strength, firm to
stiff (Su ~ 50 kPa).

Alluvial (black No formal compaction. (By Normally consolidated (low


Waghad CH/MH 63 to 93 - 35 80 to 85% 6% wet - - - - - - - - -
cotton soil) animal ?) preconsolidation pressure)

100.2% 0.1% wet Broad variation in relative compaction


Lake CL - sandy 16 (3 to Suspect rolled, well
Glacial, till 67 unk (S Dev = (S Dev = (89 to 117%) and moisture content - - - - - - - - -
Shelbyville clay 35) compacted
3.7%) 1.1%) (2.5% dry to 4.5% wet)

Post construction investigation


3
Scout 17.3 kN/m Suspect rolled, well indicated fill of stiff to very stiff
residual CL - silty
Reservation 70 to 90 unk 8 to 21 (dry unit 16 to 20% compacted (given period of strength consistency and generally - - - - - - - - -
sedimentary soils clays
Dam weight) construction) unsaturated (with some zones close
to saturation)

< 95% up to 4% Poorer compaction of lower section,


Narrow puddle core placed well wet of No formal compaction.
Glacial - Boulder Glacial - Boulder (lower), wet (lower), the upper portion of which was well
Muirhead CL/CH unk unk 25 unk wet OMC puddled OMC. Low undrained strength (Su ~ CL/SC/GC 45 20 25 Spread and compacted by
Clay. Clay. 97% 1% wet wet of OMC and of firm consistency
20 kPa) bulldozers.
(upper) (upper) (Su ~ 35 kPa).

Rolled, well compacted.


4.5% wet (4 Sandstone and Softened significantly when wetted,
Dam FC10 Alluvial CH 84 35 25 94.3% sheepsfoot roller in 150 mm Core compacted well wet of OMC SC/SM 25 6 unk unk unk Suspect poor compaction
to 8%) Weathered shale suspect poorly compacted.
layers

Initially wet, Broad zone of core material upstream


Suspect involved formal Suspect placed with formal Formed zone in upstream shoulder
Acu Dam Alluvial CH (?) unk unk unk unk then close of centreline, also as a blanket under Alluvial SC/GC unk unk unk unk unk
compaction compaction. and also downstream of centreline.
to OMC upstream shoulder.
Appendix D Page D5

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 3 of 10)
PROPERTIES OF FOUNDATION SLIDE LOCATION AND GEOMETRY
Foundation (General) Foundation (if Foundation involved in Slide)
Name Slope
General Comments on Design and Slide Classification
ASCS % Moisture Failure Slide Location State
Compaction (Hutchinson, 1988)
Soil / Rock Description Description n CF (%) PI (%) Density Comments Geometry
Class fines Content

Otter Brook Glacial till (SM/SC with gravel and Lateral bulging, not a Bulging of upstream slope, some settlement Lateral
3 m thick outer sheet of gravel and rockfill. Soil - - - - - - - - Type 1
Dam boulders) over bedrock failure. of crest. bulging

Section shows zoned earthfill embankment,


however, same materials used in core and Rock (?), Lateral bulging, not a Bulging of downstream slope. Confined to Lateral
Truscott Dam unk - - - - - - - - Type 1
outer shell. Chimney drain downstream of core weathered. failure. region of closure section. bulging
zone.

Section shows zoned earthfill embankment, normally consolidated clay, 12 m Bulging of downstream slope. Confined to
however, same materials used in core and Soil (valley thick. This was removed to Lateral bulging, not a region of closure section. Suspected lateral Lateral
Skiatook Dam - - - - - - - - Type 1
outer shell. Chimney drain downstream of core floor) bedrock in closure section with failure. movement within softer fill zone near base of bulging
zone. compacted fill. closure section.

Homogeneous embankment with upstream Shallow layer of stream bed


Rotational (?), possibly Lateral bulging of downstream (and
Troneras Dam cutoff to residual soil. High rate of construction Soil alluvium overlying deep residual - - - - - - - - Type 1 first time
only lateral bulging. upstream ?) slope incorporating crest.
to complete construction in dry season. soil profile.

Homogeneous embankment with upstream


rockfill toe. In narrow gully region embankment Bedrock in gully region. Deep Downstream slope, extended several metres
Punchina
on bedrock, slopes on deep residual soil Rock residual soil profile on abutment - - - - - - - - Rotational (?) Type 1 back into crest level at time that movement first time
Cofferdam
profile. High rate of construction to complete slopes. occurred.
construction in dry season.

Homogeneous embankment constructed of


Rotational (possibly part Failure of downstream slope within zone
Waghad CH/MH material. Placed wet without any Soil Murum over weathered bedrock - - - - - - - - Type 2 (?) first time
translational) under remediation.
formal compaction.

Homogeneous earthfill embankment with Failure in upstream slope. Extended back to


Lake embankment filter zone. Closure section of Alluvial - outwash sand of 3 m Recent alluvials stripped off to outwash Compound - translational upstream side of crest centreline. Deep
Soil - - - - - - - Type 1 first time
Shelbyville 150 m width placed from Nov 1969 to June depth over bedrock. sand layer. Central cutoff to bedrock. basal component seated. Basal plane on more plastic layer.
1970 with winter shutdown. Confined to closure section.

Relatively deep-seated slide in the upstream


Homogeneous embankment with cut-off
slope, extended back to the downstream
Scout upstream of centreline. Constructed using low Compound - translational
Residual soils and weathered edge of crest. Translational in lower portion
Reservation plasticity silty clays. Contains zones of Soil / Rock - - - - - - - - basal zone close to base Type 2 first time
bedrock. of embankment close to foundation.
Dam relatively low undrained shear strength (as low of embankment.
Movement across 75% of the embankment
as 50 kPa).
length.

Boulder clays are glacial deposits


Change in contractor in April 1941 with view to Lateral bulging of upstream slope
formed by glacial crushing of Compound - translational
increase rate of construction. Four fold incorporating crest and upstream slope.
Muirhead Soil parent rock. Medium plasticity, - - - - - - - - basal component in firm Type 1 first time
increase in rate of volume placed. Rockfill Occurred during construction when approx. 5
very stiff to hard strength earthfill zone.
zone at upstream and downstream toe. m below design crest level.
consistency.

Failure of upstream slope over significant


Broad core zone (1.5H to 1V upstream) of CH Alluvial clays, sands and gravels
Soil (valley Rotational (?), surface of width of the central portion of the
Dam FC10 clays with poorly placed random fill zone and over bedrock. Significant plastic - - - - - - - - Type 1 first time
floor) rupture not determined. embankment. Occurred shortly after end of
thin outer rock shell. clay layer of 12 m thickness.
construction.

Deep-seated failure of upstream slope


Upstream - zoned earthfill with silty clay placed Compound - translational
during construction, extended back to the
as blanket under upstream shoulder. basal component in clay
Acu Dam Soil Alluvial sands. - - - - - - - - Type 2 mid to downstream portion of the crest. first time
Downstream - zoned earth and rockfill with zone under upstream
Basal translational sliding plane on CH fill
embankment and foundation filters. shoulder.
layer under the upstream shoulder.
Appendix D Page D6

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 4 of 10)
AT AND POST-FAILURE
Slide Dimensions Post-Failure Deformation Behaviour Degree of
Name Break-up of Hydrogeology
Width, Length, Depth, D Height Slope Volume Velocity Velocity Comments
the Slide
W/L W/D Displacement (m) Comments
W (m) L (m) (m) (m) (deg) (cu.m.) (mm/day) Category Mass

0.88 m monitored
23 (no Excess pore pressures in fill during construction.
Otter Brook movement of upstream Movement only occurred when construction in progress. Once Lateral bulging of upstream slope. Suspect bulging is occurring
unk - bulging of - 21.8 - - - 40 slow Intact Piezometers indicate ru up to 0.65 to 0.8 in some
Dam face (will be greater placement suspended movement stopped within 1 to 2 days. within a wetter zone of fill.
lower slope) zones of fill. Suspect this is in the wetter zones of fill.
than this)

Significant crack in lower region of embankment (3 m above Lateral bulging of downstream slope within closure section.
0.84 m monitored
foundation). Displacement at toe of slide over-rode lower section of Suspect bulging occurred within a wetter zone of the fill. No information. Suspect potentially high excess pore
Truscott Dam 400 - - - 18.4 - - - movement of 20 slow Intact
embankment by 150 mm. At completion of construction further Shear zone at toe of slide coincided with the elevation where the pressures developed within wetter fill zones.
downstream slope.
cracking observed. filling was left exposed for several months.

Significant crack in lower region of downstream slope, commensurate Piezometer in core indicates ru in the range 0.8 to 1.0.
0.79 m, lower Lateral bulging of downstream slope within closure section.
Skiatook Dam 50 to 60 85 22 22 16 - 0.65 2.50 23 slow with identified softer and wetter zone of filling. Monitored movement Intact This is indicative of placement wet of optimum of low
downstream slope Bulging occurred within a wetter, softer zone of the fill.
of 655 mm in 29 days during which rate of construction = 0.23 m/day. permeability material.

Later part of embankment construction at a very rapid rate and


Very high excess pore pressures developed within the
Only quantitative reference to movement is that at one point the crest during the start of the wet season. Fill placed very wet (had to use
sandy silt fill, ru from 0.4 to 1 (avg of 0.75 to 0.8).
600 (suspect slow to level could not be raised due to the bulging and one night the crest tractors to compact due to bogging of rollers). After placement of
Troneras Dam 60 - - - 18.4 - - - no information Intact Very high annual rainfall (2 to 5.35 m) with only 3
high) moderate settled more than height of fill placed. Suspect that this rate may be a 1.8 m of top 9.1 m (in wet season) the fill did not rise further it
month dry season during which construction can
bit high. merely bulged. The design was changed, but further bulging and
proceed.
cracking occurred. Construction suspended when 2 m from crest.

SMPs established once movement observed. First sign was cracking Very rapid rate of construction of downstream stage of cofferdam High excess pore pressures developed within the silty
1.57 m (max) on mid on lower downstream slope and mis-alignment of surface drains. in order to complete construction before start of wet season. sand fill, ru from 0.43 to 0.7, particularly during high
Punchina
60 65 12 35 26.6 25000 0.92 5.00 to lower downstream 250 to 350 slow Embankment construction continued and further movements were Virtually intact Continued construction after first signs of movement resulted in construction rate period. Very high annual rainfall (3.1
Cofferdam
slope, 0.72 m at crest. observed. Movement stopped shortly after embankment completed ( relatively high velocities of movement of downstream slope (250 to m) with only 3 month dry season during which
1 to 2 days after). 350 mm/day). Design altered to complete construction. construction can proceed.

Moderate Excess pore pressures in fill during construction.


Suspect moderate (potentially rapid) velocity of movement given Limited information. Sections indicate movement incorporated
Waghad 61 60 18 29.5 26.5 35000 1.02 3.39 15.2 m toe, 6.4 m crest - Minor (?) Suspect pore pressures quite high given earthfill
(poss. Rapid) distance travelled upper part of upstream slope.
placed well wet of OMC.

Limited information on reservoir level operation.


Maximium stored water level in July 1971 (inundated
Rate of movement appears to have been affected by reservoir
Very slow movements (0.4 to 2 mm/day) from Oct 1970 to Dec 1971. lower half of slide). Drawdown for remedial works
level operation. Suspect a drawdown in Feb to March 1972, in
Lake 392 mm upper part 24 (max), 0.5 slow to very Virtually stopped in Dec to Feb 1972. Acceleration in movement in must have started late Jan or early Feb 1972
165 46 18 15.5 21.8 75000 3.59 9.17 Intact preparation for repairs, caused the observed acceleration in
Shelbyville upstream slope to 2 (avg) slow Feb to March of 1972, suspect due to drawdown prior to remedial (reservoir at 15 m below crest on 10 Feb 1972).
movement. Looks like lower and mid portion of slip moving before
works. Construction induced pore pressures ranged from ru of
upper portion.
0.26 to 0.64 (avg 0.4 in slide area) and showed slow
dissipation.

No water stored in impoundment up to 1988 when repairs


Scout slow First sign of movement was cracking at the end of construction. Slide Piezometers indicate variable phreatic surface within
1.8 m, scarp height at 10 reducing to undertaken. Surface of rupture determined by inclinometer and
scarp developed slowly over a period of 2 years (late 1984 to mid fill. Suspect that observations are a reflection of
Reservation 180 42 10 13.7 26.6 30000 4.29 18.00 decreasing to Virtually intact wet layers of fill. Suspect movements from 1986 onward only
crest. <1 1986). From late 1986 to 1988 movements increased during/following construction induced pore pressures and are variable
Dam very slow periods of heavy rainfall. Suspect cracks filled with water.
occurred following rainfall with no movement during relatively dry
due to moisture variation at fill placement.
periods.

High pore pressures developed in filling placed well


Movements first observed on 16 Sept 1941. Survey on 19/9/41
wet of OMC. No information on pore pressures, but
Up to 2.1 m on indicated 1.45 m lateral bulge in upstream face. Movement renewed Surface of rupture fully developed, passed through core and
Muirhead 100 62 15 15 21.8 60000 1.61 6.67 40 to 115 Slow Virtually intact suspect ru greater than 0.5 as indicated from
upstream slope. on re-commencement of filling operations (and continued for about a weaker zone of shoulder earthfill.
Knockendon Dam. Likely slow dissipation of pore
week after suspension of construction).
pressures in low permeability material.

Likely high pore pressures in core due to compaction


Movement within the embankment only (no movement at upstream
1.8 m crest, 2.1 to 2.8 Slow to Significant movements of upstream face observed during construction. well wet of optimum. Stored reservoir level was
toe). The largest movements occurred shortly after construction
Dam FC10 120 66 14 19 18.4 60000 1.82 8.57 - Slide mass accelerated after the end of construction (Jan 1937) when Virtually intact relatively high from early November 1936 to mid
m upstream slope. Moderate and suspect were related to reservoir operation. Preceeded by
cracking was observed. The movement continued for many days. January 1937. Possibllity that drawdown may have
'abnormal' movements during construction.
induced failure.

15 m movement at Slide preceeded by cracking in the crest at the downstream edge of


Failure incorporated the central part of the crest and the upstream
crest (near vertical), the core zone. This was followed by outward deformation of upstream High pore pressures developed in silty clay filling
Acu Dam 600 107 26 34.8 21.8 1200000 5.61 23.08 1.2E+06 Rapid Minor slope. Earlier failures occurred in the cofferdam (constructed of
25 m displacement at slope, for distance of 25 m, on the CH clay fill layer placed under the placed wet of OMC.
wet placed CH materials).
upstream toe upstream shoulder. Movement occurred within about 30 minutes.
Appendix D Page D7

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 5 of 10)
MECHANICS OF FAILURE
Name General Comments
Cause/s of Landsliding Trigger of Landsliding Failure Mechanism Comments on Failure Mechanics

High excess pore pressures in some piezometers reflects wetter zones of fill. These
Construction. High rate of construction zones likely to have relatively low undrained shear strength and modulus.
Otter Brook Not actually failure, only lateral bulging. Low modulus and undrained Case of lateral bulging. Authors report that factor of safety was
led to development of high pore Construction Displacements preferentially occur within the lower strength wetter soil zones. High
Dam shear strength of wetter fill zones. greater than one during construction. US Corp Dam.
pressures in wet filling zones. construction rate contributed to velocity of movement. Suspect FOS close to 1 even
though authors think otherwise.
Suspect high excess pore pressures developed in wetter zones of fill. These zones
Construction. High rate of construction Not actually failure, only lateral bulging. Low modulus and undrained likely to have relatively low undrained shear strength and modulus. Displacements Similar case to Otter Brook Dam. Lateral bulging during
Truscott Dam Construction.
within closure section. shear strength of wetter fill zones. preferentially occur within the lower strength wetter soil zones. High construction rate construction. US Corp Dam.
contributed to movement.

Construction. High rate of construction High excess pore pressures developed in wetter zones of fill. These zones likely to
Not actually failure, only lateral bulging. Low modulus and undrained have relatively low undrained shear strength and modulus. Displacements Similar case to Otter Brook Dam and Truscott Dam. Lateral
Skiatook Dam (within closure section) and high pore Construction
shear strength of wetter fill zones. preferentially occur within the lower strength wetter soil zones. High construction rate bulging during construction. US Corp Dam.
pressure development. contributed to movement.

Sandy silt placed very wet in latter part of construction. Wet placed fill of low Late start to construction required very rapid rates of
Construction. High excess pore Lateral bulging of wet placed sandy silts with low undrained shear undrained shear strength and modulus. Lateral bulging occurred once factor of safety construction to complete embankment before start of wet
approached one. Suspect material is dilative and strain hardening in undrained
Troneras Dam pressures in fill placed well wet of Construction. strength. Suspect material is dilative and strain hardening on season. Lateral bulging during period of very high construction
loading (at low confining pressures) and therefore expect movement to stop once
OMC. undrained shearing. rate. The design was altered when cracking occurred the day
construction stops. Factor of safety continually at one while construction in progress;
i.e., the bulging keeps it at one. following significant lateral bulging.

Silty sand placed well wet of optimum moisture content. Very rapid rate of construction Late start to construction required very rapid rates of
Deformation indicates formation of failure (rather than just lateral resulted in maintaining high pore pressures in fill. Suspect material is dilative and
Construction at very rapid rates of fills construction to complete cofferdam before start of wet season.
Punchina bulging) in relatively steep downstream slope (2H to 1V). Undrained strain hardening in undrained loading and therefore expect movement to stop once
well wet of OMC. High excess pore Construction. Movements occurred during period of very high construction
Cofferdam failure of near normally consolidated silty sand fill having low construction stops. Suspect relatively steep embankment slope influenced high rate
pressures developed in fill. of movement observed. Limit equilibrium analysis indicates factor of safety of 0.97 to
rate. The design was altered when relatively high rates of
undrained shear strength.
1.
movement measured on downstream slope.

During monsoon season of 1883 embankment was overtopped.


Undrained failure in near normally consolidated CH/MH fill placed well Suspect strain weakening in undrained loading due to strength loss associated with
Construction. Development of high Slide in 1884 was within section under re-construction following
Waghad Construction wet of OMC. Low undrained shear strength and likely strain orientation of clay particles and reduction from fully softened to residual strength on
pore pressures. over-top. Suspect high rates of construction may have
weakening in undrained loading. shearing. Low pre-consolidation pressures due to wet placement.
contributed to the failure.

Undrained failure. Surface of rupture passed through the weaker fill layers Partially saturated fills are likely to be strain hardening in undrained loading and strain Slow to very slow movement of slide, but significant difference
Construction and high pore pressures Suspect initial failure due to (which controlled the slide). Back-scarp through partially saturated that would weakening in drained loading due to dilative behaviour on shearing. Suspect reduced
observable in the deformation behaviour between a point on the
(initial movements may have occurred undrained instability. Suspect be dilative and strain weakening. Suspect movement observed at the end of undrained strength of wetter and more plastic fill layers may be controlling the slide
Lake slide area and a point off it. Slow movement could be due to
during construction but not observed). reservoir operation affected the construction was post-failure movement (i.e., surface of rupture had fully movement. Wetting up of the filling from infiltration (from inundation) would result in a
Shelbyville restraint on the lateral margins (from keying in of the closure
Drawdown triggered acceleration in velocity of movement, but formed). Reduction in rate of movement suggests stabilisation possibly due to reduction in the strength of unsaturated soils. This may potentially be a factor in the
reduction in pore pressures (dissipation). But slope still of marginal stability as acceleration in movement observed on drawdown in Feb 1972. Initiation of
section to the main fill) and from shear induced dilation of
movement. limited information.
indicated by accelerations possibly associated with drawdown. movements in lower portion of the slide suggest the toe region is controlling the slide. unsaturated soils in the backscarp. US Corp dam.

Suspect initial instability driven by weaker, wetter layers within fill near The width of the slide possibly indicates the basal sliding plane may be laterally very
Scout Construction. Development of extensive. It may have been wetted by rainfall or placed during a wet weather period. Slow to very slow failure starting at end of construction and
Construction. Latter the upstream toe; therefore undrained instability. Variability of
Reservation localised high pore pressure zones in Lab tests indicate fill is strain weakening in undrained loading at low confining movement continuing for 4 years. Likely undrained instability
movements triggered by rainfall. strength of filling would indicate shear plane probably passes through
Dam wetter embankment zones. pressures. Reactivation of movement from 1986 suggests the slope is of marginal initially, but later movements must be drained instability.
dilative soils. stability triggered by water infilled cracks.

Construction. High rate of construction Undrained instability of firm to stiff (and wet) filling. Core zone The monitoring results would indicate that, overall, the earthfill materials are not strain Case of lateral bulging due to limiting undrained stability
Muirhead led to development of high pore Construction. possibly strain weakening in undrained loading, but outer zone weakening in undrained loading as indicated by the reduction in velocity of the slide condition in firm to stiff medium plasticity sandy and gravelly
pressures in wet and firm filling zones. unlikely to be. onceconstruction is suspended. clay fill materials placed wet of optimum moisture content.

Not clear, combination of During construction, movements indicated significant


construction and possibly Suspect CH core is strain weakening in undrained loading due to orintation of clay deformations of upstream slope compared with downstream
Undrained failure. Likely low undrained strength of CH core and near particles and strength reduction from fully softened to residual strength. Likely low
Construction. Development of high drawdown induced. Strength slope. Core geometry flatter upstream (1.5H to 1V) than
Dam FC10 normally consolidated. CH core likely to be strain weakening in undrained shear strength and low pre-consolidation pressure of core due to wet
pore pressures in broad core. loss on saturation of the random downstream (1H to 1V). Continued movements were observed
undrained loading. placement. Suspect drawdown may have triggered slide due to loss of support from
fill zone may have had a water on upstream face and loss of strength in random fill layer once saturated.
for many days after initial failure observed indicating likely slow
significant influence on stability. rate of movement.

The lower portion of the CH core, including the section under the upstream slope,
Undrained instability of wet placed silty clay filling. Suspect slide were placed wet of optimum. Therefore, lower undrained strengths likely in this
Construction. Development of high material with instability initiated from this zone of the embankment. High shear Unusual embankment design with upstream cutoff to bedrock
brittleness possibly associated with development of the surface of
Acu Dam pore pressures in lower wet silty clay Construction stresses then transferred onto clayey sand zone at toe and also higher reaches of and upstream shoulder underlain by silty clay material. Wet
rupture through the outer clayey sand zone at the toe and/or through
zones. core. Once the failure initiated a significant reduction in shear strength within the placement of core material significant in failure.
the partly saturated earthfill in the upper part of the core. clayey sand at the slide toe and upper core zones may have contributed to the slide
brittleness and the rapid and relatively large run-out distance.
Appendix D Page D8

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 6 of 10)
Embankment Zoning and CONSTRUCTION
Classification Construction Embankment Dimensions
Date of Slide
Name Location Geology References
Failure Location
n Core Core Year Rate Height, He Length, Ratio Upstream Downstream
Type Class Comments
Width Type Started (mm/day) (m) Le (m) Le /He Slope Slope

Beene (1967)
Sedimentary - 43 (25.9 at Failure during construction of downstream
Emb & 3 to 4.5H to Stroman et al (1984)
Waco Dam USA, Texas Sep-61 Homog 1,1,1 c-vb CL (?) clay shales. 1960 48 failure 5500 127.9 2.5 to 3H to 1V slope when 4 m below design level.
Fndn 1V West (1962)
Faulting. section) Confined between faults in foundation.
Stroman and Karbs (1985)

Failure of upstream slope when


Sedimentary -
West Branch Emb & 22.5 (21.5 3 to 4.5H to constructing closure section (400 m width).
USA, Ohio 20-Oct-64 Homog 1,0,1 c-vb CL (?) shale and 1963 600 3020 134.2 3 to 4.5H to 1V Fetzer (1967)
Dam Fndn at failure) 1V At 4 m below design crest level when
sandstone, glacial
abnormal movement observed.

Failure of downstream slope during


North Ridge Emb & Glacial soils 18.3 (at
Canada Sep-53 Homog 1,1,1 c-vb CL 1953 260 unk - 4H to 1V 4.5H to 1V construction when approx. 3 m below Peterson et al (1957)
Dam Fndn (lacustrine) failure)
design level.

Sedimentary - Rapid failure of downstream slope during Sherard et al (1963)


Marshall Emb & 24.4 (at
USA, Kansas 20-Sep-30 Homog 2,0,0 c-vb CL (?) limestone and Mid 1930's unk 480 17.5 3H to 1V 2.7H to 1V construction. 3 m below design level when ENR (1937a, 1937c, 1938a,
Creek Dam Fndn failure)
shale failed. 1938b)

Cooling & Golder (1942)


Puddle Sedimentary Downstream shoulder failed during Skempton (1990)
Emb &
Chingford UK, Essex 30-Jul-37 core 8,0,1 - CH marine. London 1936 100 13 5600 430.8 3H to 1V 2.5H to 1V construction when 2.4 m below design Penman (1986)
Fndn
earthfill Clay level. Dounias et al (1988)
Charles and Boden (1985)

Puddle Sedimentary - 21 (avg incl. Failure in downstream slope during


UK, Sept to Nov Emb & 11.4 (at
Hollowell core 8,1,1 - CH mudstones 1936 shutdown 427 37.5 3 to 5H to 1V 2.5 to 4H to 1V construction when 1.7 m below design Kennard (1955)
Northampton 1937 Fndn failure)
earthfill (Upper Lias) periods) crest level. Limited movement.

Skempton and Vaughan


(1993)
Sedimentary, 110 (50 incl. Failure in upstream slope during
Zoned Emb & Rocke (1993)
Carsington UK 04-Jun-84 3,0,1 c-tn CH mudstone and 1982 seasonal 36 1250 34.7 3H to 1V 2.3H to 1V construction when very close to finished
Earthfill Fndn Byrd and Middleboe (1984)
sandstone shutdown) level.
Penman (1986)
Potts et al (1990)

Sedimentary -
Failure of downstream slope during
USA, Zoned CL/CH Emb & conglomerate, 35.5 (at Bowers (1928)
Lafayette Dam 17-Sep-28 3,0,0 c-tm 1927 85 380 10.7 3H to 1V 2.5 to 3H to 1V construction when approx. 6 m below
California Earthfill (?) Fndn sandstone and failure) ENR (1929a, 1929b, 1929c)
design level.
clay

Sedimentary -
Failure of upstream slope during
USA, New Zoned Emb & sandstone and 50 to 100 Catanach and McDaniel
Galisteo Dam 26-Jan-70 3,1,1 c-vb CL 1967 48.2 860 17.8 3.17H to 1V 2.5 to 3H to 1V construction (9.4 m below design crest
Mexico Earthfill Fndn shale (Mancos (staged) (1972)
level when movement started).
Shale)

Sedimentary - 53 (43
50 to 60 Lateral spreading of foundation. Only
Zoned Emb & clay-shales and when
Dam FC21 15-Oct-69 3,1,1 c-tk SC/GC 1966 (winter 1005 19.0 3.3H to 1V 2.25H to 1V aware when displacement in outlet conduit
Earthfill Fndn siltstone, glacial movement
shutdown) observed, no surface expression.
activity. observed)

Failure in upstream slope during


Park USA, Zoned Emb & 21.5 (at
Oct-81 3,1,1 c-tk CL (?) Lacustrine 1980's unk 366 15.0 2.8H to 1V 2.5H to 1V construction when 3 m below design crest USCOLD (1988)
Reservoir Wyoming Earthfill Fndn failure)
level.
Appendix D Page D9

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 7 of 10)
EMBANKMENT MATERIALS
Material 1 (Core) Material 2 (Shoulder fill)
Name
ASCS Density Moisture ASCS Density Moisture
Source % fines CF (%) PI (%) Compaction Comments Source % fines CF (%) PI (%) Compaction Comments
Classn Ratio Content Classn Ratio Content

CL (?) -
Limited information on materials.
Sedimentary, sandy clays Rolled, suspect well
Waco Dam unk unk unk unk unk Comprised sandy clays and broken - - - - - - - - -
shales and shaley compacted
up shale.
clays

bulk
No information. No information on materials. Sections
West Branch density = Rolled, suspect well
Glacial deposits CL (?) unk unk unk unk only indicate "impervious fill" and very - - - - - - - - -
Dam 20.4 compacted
or shales. stiff.
kN/cu.m.

North Ridge Rolled, suspect well Low plasticity clays of stiff to very stiff
Glacial clays CL unk unk 10 to 18 96.9% 0.1% dry - - - - - - - - -
Dam compacted strength consistency.

Rolled, well compacted.


Marshall Placed in 200 mm layers and Limited information on fill materials,
unk CL (?) unk unk unk unk 14 to 16% - - - - - - - - -
Creek Dam compacted by 12 to 14 refered to as "impervious fill".
passes of sheepsfoot roller.

No formal compaction. Select zone adjacent to core and


Marine 100 suspect
well wet of No formal compaction. Placement of puddle core. Very soft Alluvial (select and tracked by wagon, bank outer bank fill. Outer bank fill
Chingford sedimentary CH (close to high 53 unk CH very high very high up to 100 unk wet of
OMC Tracked by wagon (Su ~ 10 kPa). bank material) material spread and tracked comprised a mix of CH alluvial clay
(London Clay) it) optimum
by dozer. and ballast.

Puddle core and select zones


Sedimentary, adjacent to core using Upper Lias Sand and
puddle, presume placed in
Hollowell Upper Lias clay CH Very high 45 32 unk wet OMC clay (high plasticity). Placed wet of unk sand-rock unk unk unk unk unk unk -
thin layers.
(marine clay) OMC. Suspect of low undrained material.
shear strength.

Stiff consistency of core fill (Su ~ 60


Rolled, well compacted. 200 to 65 kPa). Strain weakening in Rolled, well compacted. 300
Periglacial and close to Heavily over-consolidated after
Carsington CH unk 56 38 93.2% 8% wet to 250 mm layers with 28 undrained loading with Sur of 30 kPa. Weathered mudstone SC/CL unk 34 22 101.10% mm layers compacted with 7
residual OMC compaction.
tonne padfoot rollers. Pre-consolidation pressure of 140 to tonne grid rollers.
150 kPa.

Weathered and Sandstone and < 15% 3 Rolled, well compacted.


50 (70% Rolled, well compacted. Materials from Orinda formation, 19.5 kN/m
argillaceous conglomerate (20% less Placed in 200 mm (upstream) Materials also from Orinda
Lafayette Dam CL/CH (?) high finer than unk unk dry of OMC Placed in 200 mm layers and relatively young (Pliocene) deposit. GC (?) low unk (bulk unk
sedimentary (conglom. only for than 5 to 300 mm (d/str) layers and formation.
5 micron) heavily rolled. Core slopes 0.5H to 1V. density)
bedrock. d/stream shoulder). micron) heavily rolled

9.3 to 15.2 to
CL - clays unk (low to 17.1 kN/m3 26.1% Limited information. Undrained unk (low to 3 unk (Spec. Similar materials used for outer
Alluvial and/or Rolled, suspect well Alluvial and/or Rolled, suspect well
Galisteo Dam and sandy high (?) unk medium (dry), Spec. (Spec. strength tests indicate firm to very CL (?) unk unk medium 19.6 kN/m 3% dry to earthfill as for core. Stiff to very stiff
residual compacted residual (dry), Spec. compacted.
clays plasticity) > 95% OMC to 2% stiff. plasticity) OMC) strength consistency.
wet) > 95%

Rolled, well compacted. sand, gravel


Limited information. Suspect dilative. Rolled, limited compacted. Downstream shoulder comprised
0.2% dry Placed in 150 mm layers and and cobbles
Dam FC21 Glacial soils SC/GC low low unk 99.8% Flat upstream slope (1.75H to 1V) Glacial soils unk unk non plastic unk unk Placed in 300 mm layer and outer zone of dumped cobbles and
(avg) compacted by tamping (cohesionles
and steep d/str slope (0.5H to 1V). compacted by crawler tractors. boulders.
rollers. s)

Park Limited information on fill materials,


lacustrine (?) CL (?) unk unk unk unk unk unk Glacial (?) sand (?) unk unk unk unk unk unk Only referred to as "sand shell"
Reservoir refered to as "impervious fill".
Appendix D Page D10

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 8 of 10)
PROPERTIES OF FOUNDATION SLIDE LOCATION AND GEOMETRY
Foundation (General) Foundation (if Foundation involved in Slide)
Name Slope
General Comments on Design and Slide Classification
ASCS % Moisture Failure Slide Location State
Compaction Soil / Rock Description Description CF (%) PI (%) Density Comments (Hutchinson, 1988)
Geometry
Classn fines Content

Clay shales very stiff to hard, horizontally


Surface of rupture along Deep-seated slide in downstream slope.
Homogeneous earthfill embankment with bedded. Extensive pre-sheared bedding
Alluvial - 2.5 to 5 m of clays and pre-sheared bentonitic Compound - translational The slide back-scarp extended back to the
Waco Dam embankment and foundation filter zone. Soil CH (?) high (?) high (?) high (?) unk unk plane in Pepper Shale between near Type 5 first time
gravels over shale bedrock. clay shale bedding basal component. upstream slope and the slide toe was 140 m
Central cut-off to bedrock foundation. vertical faults. Limited lateral resistance
defect. beyond the downstream toe of embankment.
along faults.

Glacial lake deposits overlain by Slide in upstream slope of embankment


Homogeneous embankment with downstream alluvial silty sands. Soft to stiff clay 14.8 21 to 40%
Compound - translational extending back to mid to upper level of
West Branch drainage blanket. Central cut-off to clay stratum (15 to 18 m thickness) at 5 to Soft to stiff clay stratum, CL - silty kN/cu.m. (several % Low to medium plasticity silty clay, soft to
Soil high unk 9 to 28 basal component in clay Type 5 foundation. Limited to zone of closure first time
Dam foundation. Closure section about 400 m 8 m depth below silty sands. laminated (Su > 19 kPa) clay (dry below stiff.
foundation. section. Basal surface 4 to 8 m below
width. Underlain by dense silty sands and density) liquid limit)
very stiff clays.
ground surface.

Homogeneous embankment with central cut-off 3 to 6 m of silty fine sand 15.3 Of firm to stiff strength consistency. Compound (?) - with
Slide in downstream slope during
North Ridge to clay foundation. Embankment filter in overlying firm to stiff high kN/cu.m. Undrained creep tests indicate Su ~ 40 translational basal
Soil High plasticity clays CH high (?) high (?) 51 36.50% Type 5 construction. Cracks observed on upper first time
Dam downstream shoulder and filter on foundation. plasticity clays. Assume clays (dry kPa (compared with 55 kPa from quick component in foundation
portion of upstream slope. Deep seated.
Constructed using low plasticity clays. are lacustrine glacial deposits. density) triaxial tests) clays.

Earthfill embankment with well compacted Silty clay layer, 3 to 6 m 30 to 60% Alluvial foundation comprises a mix of
15 m thickness of alluvium (clay, Compound (?) - with
Marshall central and upstream clay zones. Downstream thickness, soft to firm very (40 to 80% soil types. Not sure how extensive this Deep-seated slide in the downstream slope.
Soil sandy clays silts and sands) over CL/CH (?) unk unk unk translational basal Type 5 first time
Creek Dam section comprises "loose rock and earth" rolled (not sure how extensive high finer than layer is. References also indicate Slide crest at the upstream edge of the crest.
shale and limestone bedrock component in foundation.
in 450 mm layers. Rockfill zone at toe. layer is) 5 micron) presence of loessic soils.

Suspect select and bank material placed


Alluvial deposits comprising soft Deep-seated slide in downstream shoulder
relatively wet given low undrained strengths Clay (up to 4 m thick), Compound - translational
peaty clays and clays overlying very up to 14.9 kN/m3 Upper clay zone very soft to soft, very during construction. Back-scarp at
Chingford presribed to these materials (15 to 28 kPa). Soil soft to very soft (Su ~ CH very high 90% basal component in soft Type 2 first time
river gravels. Up to 9 m depth the high 100 (bulk ?) high plasticity and high moisture content. embankment crest, on the upstream side of
Ballast filter and rockfill toe downstream of 14 kPa) clay foundation.
London Clay. the select fill.
puddle core.

Puddle core earthfill dam. Drains installed in


Alluvial or peri-glacial (?), layer of Compound - translational Deep-seated slide in downstream shoulder.
lower downstream zone comprising vertical No details on clay except that of soft to
Hollowell Soil soft to very soft clay 1.5 m depth Soft to very soft clay unk unk unk unk unk unk basal component Type 5 Back-scarp extended back to upstream side first time
rubble drains (perpendicular to centreline) and very soft undrained strength.
below foundation strip level. rotational backscarp. of select zone (on the crest).
rockfill toe.

Heavily over-consolidated
Failure in upstream shoulder, extended back
Unusual boot design to upstream side of core. Peri-glacial high (preconsolidation pressure > 600 kPa).
Compound - translational to downstream side of core. Initial failure
Core of CH clay placed well wet of OMC. Soil (rock in Periglacial high plasticity clay plasticity clay with Significantly reduced peak strength on
basal component in confined to 110 m width but spread to about
Carsington Drainage layers placed in outer embankment gully overlying residual soils and solifluction shears CH unk 62 43 unk 38% solifluction shears (oriented parallel to Type 2 first time
foundation, rotational 500 m. Suspect initiated in zone around
zones. Small berm constructed at upstream section) weathered bedrock profile. (reduced strength on ground surface) and small strain required
through core. boot where FEA indicated high shear
toe over deepect section of embankment. shears). to residual strength. Strain weakening in
stresses.
drained loading.

Thick, near horizontal sandy clay stratum


Zoned earthfill embankment with thin core Alluvial (clays, sands and gravels)
overlain by up to 5 m of surficial soils. Compound - translational Deep-seated slide in downstream slope, 12
(0.5H to 1V slopes) and broad sandy and in broad valley section (2 - 3 m uo Sandy clay alluvium CH - fine unk (but
Lafayette Dam Soil high high (?) high (?) unk Confined to deeper alluvial profile in basal component in clay Type 5 to 15 m depth below ground surface. Back- first time
gravelly (GC) outer earthfill zones. Shallow tp 20 m thick). Bedrock on (fine sand). sandy clay saturated)
broad gully region. No indication of (alluvium) foundation. scarp at upper edge of upstream slope.
central cut-off then sheet piling to bedrock. abutment slopes.
shear strength but suspect relatively low.

Zoned earthfill embankment with thick core (2H Deep-seated slide in upstream slope, limited
Alluvium (clay, sand, gravel and Clay stratum in alluvial 3 Compound - basal
to 1V upstream and 1H to 1V downstream). 15.6 kN/m Laboratory testing indicates of firm to stiff to broad gully region of embankment. Slide
silty clay) up to 24 m thickness in foundation. Up to 5 m translational component
Galisteo Dam Thin outer earthfill zone on upstream side. Soil CL/CH (?) unk unk unk (dry 24% strength consistency (Su ~ 50 kPa), with Type 5 back-scarp at upstream side of crest. Basal first time
broad gully. Abutments on thick and of firm to stiff on weak clay foundation
Similar materials used for all zones, but wetter density) very stiff layers (?). zone about 8 to 10 m below ground surface
bedrock. strength consistency. layer.
moisture spec. for core. level.

Glacial soils of shallow depth in


Zoned earthfill embankment. Core has Glacially disturbed Upper glacially disturbed bedrock zone of Compound - basal
gully region overlying shale Lateral spreading, both upstream and
relatively flat upstream slope with relatively thin upper bedrock zone. 0.5 to 3 m thickness comprising CH clay translational movement Lateral
Dam FC21 Soil bedrock. Upper zone of shale CH high high high unk unk Type 5 downstream, in central section of
outer upstream zone. Central cutoff to Contains slickensided with some cobbles and slickensided on pre-sheared spreading
glacially disturbed and contains embankment. Greater upstream movement.
bedrock. bentonitic shear zones. shear surfaces. bentonitic clays.
bentonitic clay layers.

Lacustrine varved clays and silts


Zoned earthfill embankment with thick central Compound - translational Slide in upstream slope with back-scarp at
Park under upstream shoulder, Glacial Lacustrine varved clays Limited information on properties of
core (1H to 1V slopes) and outer sand shells. Soil CL (?) high (?) unk unk unk unk basal component in Type 2 (?) the upstream edge of the crest. first time
Reservoir morraine under central and and silts lacustrine deposits.
Chimney filter on downstream side of core. lacustrine foundation. Translational within foundation.
downstream shoulder.
Appendix D Page D11

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 9 of 10)
AT AND POST-FAILURE
Slide Dimensions Post-Failure Deformation Behaviour Degree of
Name Break-up of
Comments
Hydrogeology
Width, Length, Depth, D Height Slope Volume Velocity Velocity the Slide
W/L W/D Displacement (m) Comments
W (m) L (m) (m) (m) (deg) (cu.m.) (mm/day) Category Mass

Zone of bulging (toe of slide) located 140 m downstream of the


Slide first noted when cracks observed on the upper part of the Relatively Very high excess pore pressures (r u of 0.7 to 1)
7.9 m (max), embankment toe. Suspect slow acceleration of movement over
1600 to 1700 downstream slope on 4/10/1961. Initial signs of movement dated minor, mostly in developed on the pre-sheared bentonitic seams and
more than 30 days associated with restraint on lateral margins
Waco Dam 210 260 40 26 19.7 1.0E+06 0.81 5.25 downstream shoulder. moderate back to 17/9/61. Movement accelerated slowly over many days (at faults in the foundation. Excess construction pore
(max) rear portion of through the embankment. Once a shear surface fully developed
7.1 m crest settlement. least 30 days) then accelerated rapidly to 1700 mm/day by 25/10/61 pressures were measured as far as 800 from the dam
before slowing.
slide and at toe. then the slide accelerated. Also passive restraint in toe region of
on the pre-sheared surfaces.
slide a factor.

Joint separation (outlet conduit) between sections 15 and 16 first Only indication of movement was spreading of outlet conduit (i.e. High pore pressures in soft clay stratum in the
0.76 m (separation of noted on 20/10/64. Filling continued whilst movement in outlet conduit no cracking observed on embankment). Fill placement stopped on foundation, r u of 0.8 to 0.85. Maybe less than this
West Branch conduit joints, 220 mm 21 slow Virtually intact
200 150 27 21.5 14 300000 1.33 7.41 monitored. Velocity of movement increased during fill placement. 14/11/64. Cracking was observed on the embankment slopes on given suspected artesian pressures in underlying
Dam between 15 and 16) When filling was suspended deformation continued at a near constant 19/11/64. The crest was then cut back and the movement sands and gravels of 60 kPa. Excess pore pressures
rate. The crest was cut back to stp the movement. stopped. very slow to dissipate.

1.37 m, mid No indication of velocity of movement. Cracking was observed on the Limited information on movement. Suspect driven by weak
High excess pore pressures developed within the clay
North Ridge downstream slope, 0.9 slow to upper upstream slope and mid height on downstream slope (no foundation and therefore movements most apparent in toe region.
unk 110 18 18.3 12.5 - - - unk Virtually intact foundation, particularly in the upper layers (ru approx.
Dam m overthrust beyond moderate vertical displacement). The downstream toe and for 20 m beyond Amount of movement and relatively flat slopes would suggest
0.7 to 0.85).
toe. heaved and moved outward 0.9 m. movement probably occurred at a sloe to moderate velocity.

15 m at crest and on Crack initially observed near crest of embankment and then observed
Secondary crack observed on upper part of downstream slope No information. Suspect potentially high excess pore
Marshall to increase in width. Movements accelerated, with the major portion
240 80 26 24.4 20.5 300000 3.00 9.23 mid downstream 1.5E+06 rapid Some break up typical of compound type slides with basal translational pressures developed within lower strength, low
Creek Dam of the movement occurring in 15 to 20 minutes (avg rate of 1 m/min).
slope. component. permeability clay zones in foundation.
Heave movements (up to 4.3 m) for distance of 40 m beyond toe.

Suspect high pore pressures in soft clay foundation


Movement evident several days prior to slide (broken water pipe at Deep seated compound failure with translational basal component
0.6 m at crest, 4.3 m at and embankment. Materials would be of very low
Chingford 90 33 8.4 7.9 21.8 13000 2.73 10.71 - Moderate toe, repaired then broke again). Suspect failure movement was at Virtually intact through soft foundation and rear rotational backscarp through
downstream toe. permeability due to their high plasticity and high fines
moderate rate ("after a while movement ceased"). select and puddle core zones.
content.

80000 Limited information on movement. Translational within soft to very Week of wet weather followed 33 day period of dry
(main 490 mm at crest, 445 soft foundation, rotational backscarp within select and core of weather, during which a significant amount of fill was
Hollowell 200 50 11 11.4 14 4.00 18.18 - slow Suspect slow given short distance of displacement. Intact
portion mm at toe embankment. Suspect significant lateral restraint due to limited placed. Suspect high pore pressures in core and
12500) persistence of the weak foundation. select fill, and possibly in clay foundation.

Heavy rain on weekend that slide initiated. High pore


IVM in core indicated shear movements developing in October 1983.
Upstream slide considered to have initiated in boot region where pressures in core (ru ~ 0.85) once pre-consolidation
Disproportionate amount of lateral movement of upstream slope in late
11.1 m at crest, 18.1 m 17000 to high shear stresses present. Significant restraint on lateral pressure exceeded (140 to 150 kPa). No excess pore
Carsington 500 105 22 28 18.4 550000 4.76 22.73 Moderate May 1984. Major slide started on 2nd to 3rd June and took place over Limited
at upstream toe. 24000 margins resulted in slowly accelerating movements over several pressures in foundation or Zone 2 outer fill (some in
5 to 6 days. Initially at slow rate (< 5 mm/hr) and slowly increased
days before major movement on night of 6 June. Zone 1 outer fill adjacent to core). Slow dissipation of
over several days. Peak rate over-night on 5 June.
pore pressures in core.

Cracks observed at crest and along downstream toe (17/9/28 ?) Toe virtually No monitoring. Suspect movement similar to Carsington and
Water table at 6 to 9 m depth below ground surface.
followed by major movement of embankment (about 6 m) over 3 days. intact, moderate Waco in that restraint on lateral margins resulted in a slow initial
7 to 10 m at crest, 9 m 2000 (avg High pore pressures only encountered in clay stratum
Lafayette Dam 160 155 50 35.5 20 700000 1.03 3.20 Moderate Bulging at toe preceeded major movement of embankment. to large degree phase of acceleration before the slide was released at relatively
at downstream toe. over 3 days) in foundation (r u approx. 0.53 to 0.65). Reservoir level
Movement continued at a decreasing rate for a further 8 days after of break up of high velocity. Significant period of continuing movement.
very low when failure occurred.
major movement. backscarps. Numerous deep cracks observed in crest region.

Movement first observed on 26/1/70 when cracks observed on mid A significant amount of movement had occurred before the slide
Query over piezometer results. Excess pore
1.4 to 2 m (max) on upstream slope. Survey indicated up to 0.9 m movement on lower was detected, and is indicative of initiation of the slide in the
pressures in core up to ru values of 0.65 to 0.74. No
Galisteo Dam 400 150 38 38.5 17.5 1.2E+06 2.67 10.53 mid to lower upstream 50 (max) slow slopes. Placement suspended and toe berms added but movement Virtually intact foundation. Movements stopped once construction completed.
piezometers in foundation, but suspect high pore
slope. continued when construction continued. Series of berm additions and Movement of the upstream slope was significantly greater than
pressures in weak clay foundation stratum.
stoppages to completion. Total displacement up to 2 m. that recorded for downstream slope (9 times greater).

less than 1 Unaware of movement until joint movements observed in outlet


0.9 m lateral conduit on 14/10/69 (max 835 mm). Inclinometers installed after this No surface expression of movement. This may have been Very high pore pressures recorded at till shale
(1970
Dam FC21 unk 170 42 43 18.4 unk - - displacement slow (?) recorded 50 mm displacement during 1970 construction period at Intact obscured by construction, but suspect would have observed interface during construction (r u of 0.9 to 1). Very
construction
(upstream) of conduit. rates < 1 mm/day. Rate of movement likely to have been much cracking over shutdown period. slow dissipation of pressures.
period) quicker during earlier un-recorded stage of movement.

Slide occurred overnight at suspected slow to moderate velocity given


Park slow to Limited information. Section indicates movement in upstream
unk 80 17 21.3 19.7 - - - 0.9 m unk limited deformation. No reference to residual movements over Virtually intact Excess pore pressures within varved clays and silts.
Reservoir moderate zone of core and translational within foundation.
subsequent days.
Appendix D Page D12

Table D1.1: Case studies of slope instability of embankment dams – slides during construction (Sheet 10 of 10)
MECHANICS OF FAILURE
Name General Comments
Cause/s of Landsliding Trigger of Landsliding Failure Mechanism Comments on Failure Mechanics

Undrained failure at residual strength along pre-sheared bedding Translational movement on pre-sheared bentonitic bedding
Construction. High excess pore plane in foundation. Brittleness of slide mechanism associated with Approximately 1 m of deformation occurred before slide accelerated at a rapid rate. plane in Pepper Shale foundation. Slide limited to region
Likely limited lateral restraint on near vertical faults. Interesting that toe of slide
Waco Dam pressures on pre-sheared bedding Construction. shearing through dilative, partially saturated embankment materials in between fault zones where Pepper Shale exposed at ground
located 140 m downstream from dam, indicating this was the nearest positively
surface and faults in foundation. backscarp and lateral margins, and also passive resistance at the toe oriented weakness in the foundation.
surface. Either side of faults no movement and much lower
of the slide. foundation pore pressures were observed.

Possible strain weakening of the overall slide mass due to restraint on lateral margins Case of lateral spreading of foundation during construction.
Construction. High excess pore and backscarp through the partially saturated embankment fill. Suspect the low Total movements relatively small, but continued movements (at
West Branch
pressures developed in soft clay Construction. Undrained failure along soft clay foundation. plasticity, soft clay foundation is not strain weakening in undrained loading. Overall, near constant rate) once construction suspended indicate likely
Dam
foundation. the velocity of movement was slow and indicative of limited strain weakening to that strain weakening of the slide mass. Movement only stopped
point. when crest level cut back some 4 m. US Corp dam.

Undrained failure in low strength (firm), high plasticity clay foundation. Failure controlled by low undrained strength clay stratum in foundation, Not clear if foundation is lacustrine deposits, but suspect so.
Construction. Development of high Undrained strain weakening in CH clay foundation would contribute to distance along which translational movements took place. Suspect some degree of
North Ridge High rate of construction contributed to failure as pore
pore pressures in low strength clay Construction. of movement. Suspect brittleness of slide mechanism associated with
Dam strain weakening of both the foundation (in undrained loading) and within the pressures in foundation had no time to dissipate and therefore
foundation. development of shear surface through stiff to very stiff partially saturated
embankment fill.
embankment fill. strengthen the foundation.

Construction. Trigger for rapid Suspect initial movements may have been associated with undrained Consider flow liquefaction of contractant loessic soils to be a possible cause Interesting failure given the large volume involved and the rapid
Construction. Likely development of
Marshall movement may have been instability of low strength clay layers in the foundation. However, this for the rapid movement of the embankment. Liquefaction may have been rate of failure. Suspect flow liquefaction within loessic soils in
high pore pressures in low strength
Creek Dam collapse and flow liquefaction of would not account for the rapid rate of movement of such a large triggered by movements associated with undrained instability of low strength the foundation may be a possible reason for the rapid
clay foundation layers.
loessic soils. volume. Suspect flow liquefaction of loessic soils. clay layers. movement.

Construction. Development of high Undrained failure. Near normally consolidated CH earthfill and Finite element analysis by Dounias et al (1988) was able to
pore pressures. Relatively rapid rate of foundation. Both materials likely to be strain weakening in undrained Strain weakening in undrained loading associated with orientation of clay particles and reasonably replicate the failure by total stress modelling. Very
Chingford Construction. reduction in strength from fully softened to residual undrained strength on shearing.
construction for puddle core earthfill loading, hence contributing to the distance and likely velocity of this high shear stresses at the embankment toe and in the puddle
Normally consolidated.
dam. small volume slide. core.

Construction and pore pressure Likely low undrained shear strength of foundation clay and embankment core and
select fill. All likely near normally consolidated at relatively low confining pressures. Further movements observed after re-commenced construction
development. Rainfall possible led to
Hollowell Construction Likely undrained failure through foundation and core zone. Potential undrained strain weakening of CH core and possibly foundation. Suspect (with new toe berm in place). Suspect lateral restrain
increased bulk density of the
significant lateral restraint on the margins, due to the limited persistence of the weak significantly limited rate and amount of movement.
permeable shoulder fill. foundation, limited the amount of movement.

Finite element analyses by Potts et al (1990) indicates concentration of shear stresses


Progressive failure. Undrained failure through wet CH core, which lab and strains in core and boot region. Suspect undrained strength of core exceeded in Classic case of progressive failure. CH core is strain
Construction. High pore pressures in tests indicate is strain weakening in undrained loading. Drained failure boot region resulting in plastic straining and transfer of load onto foundation and other weakening in undrained loading and foundation is strain
Carsington Construction. areas. Shear capacity of solifluction shears also exceeded which transferred load to
core. Solifluction shears in foundation. through foundation. Solifluction shears in foundation contributed weakening in drained loading. Presence of solifluction shears
ends of shears. Further straining resulted in strain weakening of core (undrained) and
significantly to progressive failure mechanism. foundation (drained). Eventually factor of safety at or less than 1 resulting in
significantly contributed to progressive failure mechanism.
acceleration of slide mass.

Significant restraint on the lateral margins due to extension of the slide


Undrained failure within clay foundation. Back-scarp through high undrained beyond the possible extent of weak clay alluvium in the foundation. This
Construction. High pore pressures strength and partially saturated embankment materials. Suspect slide No monitoring. Slide occurred at a relatively "rapid" velocity.
would have given brittleness to the slide in addition to that associated with
Lafayette Dam developed in low strength alluvial clay Construction. brittleness associated with development of back-scarp through partially Suspect early warning signs of the slide were not detected until
shearing through the dilative embankment materials. Bulging in downstream
stratum in foundation. saturated soils and significant restraint on the lateral margins (limited it was too late.
persistence of weak foundation).
toe region prior to movement of embankment indicating that the foundation
was driving the slide.
Suspect weak foundation layer is near normally consolidated under the embankment Would appear that the toe of the slide did not move in unison
Construction. Development of high Undrained failure along weak, low undrained strength clay layer in the
load. Relatively high pore pressures in core possibly suggest undrained failure over a width of 400m. The initial 0.9 m displacement at the toe
excess pore pressures in clay foundation. Further movements after berm additions indicates strain
through core (but of variable strength so potentially some dilation and strain
Galisteo Dam Construction. weakening of slide mechanism. Strain weakening possibly associated with was limited to a width of about 60 m. Also latter movements
foundation and undrained strength (of weakening associated with shearing through core). Suspect significant restraint on
backscarp development through partially saturated earthfill and/or restraint on
the lateral margins due to tapering out of clay stratum at abutments and lateral
varied significantly across the width of the embankment within
this layer) exceeded. the lateral margins.
shearing through embankment.
the broad gully region.

Suspect that a surface of rupture through the embankment had not fully developed.
Lateral spreading (translational movement) along pre-sheared bentonitic clay
The slide initiated in the pre-sheared foundation (exceeded shear capacity) but the Case of lateral spreading due to translational movements along
Construction. Pre-sheared bentonitic seams at residual strength. Shear stress on basal shear plane exceeded,
Dam FC21 Construction. shear stress capacity of of the earthfill had not yet been exceeded. Possible passive pre-sheared bentonitic clays in the foundation. No surface
clay zones at residual strength. particularly in region close to embankment centreline. Consider that a slide
resistance to sliding on shear plane in toe region. Also, restraint on the lateral expression of movement (cracking) observed.
surface had not fully developed through the embankment.
margins (much greater glacial soil cover over bedrock on abutments).

Construction. Likely development of Delay in failure suggests progressive, strain weakening failure.
Park Possible undrained strain weakening within foundation. No indication of Very limited information on failure and also embankment and
high pore pressures in foundation, not Construction. Possibly undrained failure within foundation. Strain weakening maybe
Reservoir width, but possible restraint on lateral margins may have delayed the failure. foundation properties.
sure about core (no information). associated with development of shear surface through core.
Appendix D Page D13

Table D1.2: Case studies of slope instability of embankment dams – slides in the upstream slope during drawdown.
Embankment Zoning and EMBANKMENT MATERIALS
Years Since Embankment Dimensions
Date of Classification Slide Material 1 (Core)
Name Location End of Geology
Failure Location
Construction n Core Core Height, He Length, Ratio Upstream Downstream ASCS Density Moisture
Zoning Class Le (m) Le /He Comments Source n % fines CF (%) PI (%) Compaction Comments
Width Type (m) Slope Slope Class Ratio Content

Embankment height reduced and downstream Alluvial (black (By animal ?). Not formally Near normally consolidated (low
Waghad India 1919 34 Homog 2,0,0 c-vb CH/MH Emb unk 26.9 1270 47.2 3H to 1V 4H to 1V CH/MH 63 to 93 - 35 80 to 85% 6% wet
slope flattened after 1884 failure. cotton soil) compacted. preconsolidation pressure)

Zoned earthfill embankment in main section. In


2.5H to 1V Rolled, well compacted. Well compacted, low plasticity, sandy clay.
Sedimentary - 42 (28.7 region of the failure (left abutment) the CL (some ML) 100.5%
(upper 25 m), 1% dry OMC Placed in 225 mm layers and Likely high undrained shear strength and low
Dam FD2 1993 32 Homog 0,0,0 c-vb CL Emb shales and failure 609 14.5 2 to 4H to 1V embankment is homogeneous with a 2H to 1V Alluvium - low plasticity 81 28 15 (S Dev = 10
10H to 1V (S Dev = 1%) compacted by 12 passes of a permeability (2 x 10- m/sec). Variability of
sandstone section) upstream slope, and constructed on an upstream sandy clay 2.3%)
below this. sheepsfoot roller. material plasticity and grading.
sloping hillslope (7 to 13 degrees).

3 (Spec 3% dry to Rolled, well compacted.


2H to 1V (5H Homogeneous earthfill embankment with drain and 15.2 kN/m 3% wet). In 1933 -
Sedimentary - Alluvial / CL/CH - silty Placed in 150 mm layers, Described after 1980's investigation as "moist
Dam FD3 2-Aug-31 20 Homog 2,0,0 c-vb CL/CH Emb 27.4 1980 72.3 to 1V lower 2.1H to 1V small rockfill zone at downstream toe. Upstream 85 to 100 30 to 45 14 to 34 (dry)
shale Colluvial clay 17 to 24% (avg compacted 12 to 21 tonne and firm with some wet and soft zones"
slope) faced with concrete panels. 95 to 100% 1.5% wet) tamping rollers
Lino (1997) considers filling to have been
Residual / over-compacted. Compacted with heavy
Homogeneous earthfill embankment with central Rolled, well compacted. Cat
Mondely Dam France Oct-81 1 Homog 1,1,1 c-vb CL Emb unk 24 unk - 3H to 1V 2.5H to 1V Weathered CL 80 high (?) 10 to 30 98% Placed wet of OMC rollers at moisture contents wet of optimum.
chimney filter and foundation filter. 825 padfoot rollers
argillaceous Low undrained shear strength, firm to stiff (Su
~ 40 to 80 kPa).

21.3 (17.8
April 1961 Zoned Zoned earthfill embankment with central cut-off and Alluvium (black
Sampna Tank India 4 and 7 3,0,0 c-tk CH Emb unk failure unk - 2H to 1V 2 to 3H to 1V CH high high high unk unk unk No information.
and 1964 Earthfill rockfill downstream toe. cotton soils)
section)

Puddle core earthfill dam constructed in 1866. Place in 75 mm layers, Finer material from colluvial borrow area.
USA, Puddle core Sedimentary -
Pilarcitos Dam 1969 103 8,0,0 - CL (?) Emb 29 158 5.4 2.5H to 1V 2H to 1V Downstream raise in 1874 (6 m raise) with puddle Colluvial CL (?) > 50 unk > 11 unk unk moistened then tracked by Presume placed wet of OMC, has low shear
California earthfill sandstone (?)
clay extending up the upstream face. hauling wagons and carts strength and low preconsolidation pressure.

Homogeneous embankment with very steep Rolled, reasonably Medium plasticity, slightly sandy, silty clay.
fluvial and/or CL - sandy
Charmes Dam France 20-Oct-09 3 Homog 0,0,0 c-vb CL Emb Sedimentary 16.8 365 21.7 1.5H to 1V 1.9H to 1V upstream and downstream slopes. Concrete slab 80 10 18 unk unk compacted. Placed in thin Some formal compaction, but suspect limited
residual silty clay
facing upstream slope. Constructed 1902-06. layers and rolled with rollers. given period of construction.

Suspect rolled, well -8


Homogeneous embankment with foundation filter CL/CH (?) - Limited information. Permeability of 10
Enista Dam Bulgaria 1982 13 Homog 1,0,1 c-vb CL/CH (?) Emb unk 32.2 1370 42.5 3H to 1V 2.75 to 3H to 1V alluvial (?) unk unk unk unk unk compacted (given period of
under downstream shoulder. silty clays m/sec.
construction)

Suspect rolled, well -9


Homogeneous embankment with no filters. Limited information. Permeability of 1.5 x 10
Telish Dam Bulgaria 1982 14 Homog 0,0,0 c-vb CL (?) Emb unk 29.5 1960 66.4 3H to 1V 2.5 to 3H to 1V alluvial (?) CL (?) unk unk unk unk unk compacted (given period of
Constructed on an alluvial foundation. m/sec.
construction)

Suspect rolled, well


Homogeneous embankment with rockfill zone at CL (?) - silty Limited information. Silty clay of low
Drenovets Dam Bulgaria 1984 16 Homog 2,0,0 c-vb CL (?) Emb unk 29 unk - 3H to 1V 2 to 2.5H to 1V alluvial (?) unk unk unk unk unk compacted (given period of -9
downstream toe. clays permeability, 2 x 10 m/sec.
construction)
Suspect rolled, well No information. Constructed of sandy silty
unk (suspect Homogeneous embankment. Little information CL (?) - sandy -9
Mechka Dam Bulgaria 1981 11 Homog unk c-vb CL (?) Emb unk 32 unk - unk alluvial (?) unk unk unk unk unk compacted (given period of clays, suspect of low permeability, 2 x 10
3H to 1V) available. silty clays
construction) m/sec.
Suspect rolled, well No information. Constructed of sandy silty
unk (suspect Homogeneous embankment. Little information CL (?) - sandy -9
Sushitsa Dam Bulgaria 1981 11 Homog unk c-vb CL (?) Emb unk 32.5 unk - unk alluvial (?) unk unk unk unk unk compacted (given period of clays, suspect of low permeability, 2 x 10
3H to 1V) available. silty clays
construction) m/sec.

Suspect rolled, well


Malko Sharkovo Zoned Zoned earthfill embankment. Very limited
Bulgaria unk 1 3,0,1 unk CL (?) Emb unk 41 1050 25.6 unk unk unk unk unk unk unk unk unk compacted (given period of No information.
Dam Earthfill information and no section.
construction)

Medium plasticity silty clays. Possibly formal


Homogeneous earthfill embankment with steep
compaction, but suspect limited given period
Wassy Dam France 1883 1 Homog 0,0,0 c-vb CL Emb unk 16.5 unk - 1.5H to 1V 1.9H to 1V upstream and downstream slope. Constructed unk CL - silty clay 85 < 10 18 unk unk unk
of construction. Calcareous material in some
1881 to 1883. Brick facing of upstream slope.
samples.

No formal compaction.
Mount Pisgah USA, unk (possibly Homogeneous earthfill embankment with steep Decomposed Probable very poor compaction of low
5-Sep-28 18 Homog 0,0,0 c-vb ML Emb 23.2 195 8.4 1.5H to 1V 2H to 1V ML - sandy silt 55 10 to 15 10 unk unk Probable very poor
Dam Colorado granitic) upstream and downstream slopes. granite plasticity sandy silt earthfill
construction practice.

Likely limited level of compaction with high


No formal compaction.
Homogeneous earthfill embankment with steep degree of variability. Lower portion of fill
Alluvial / Spread in thin layers,
Forsythe Dam USA, Utah 1921 1 Homog 0,0,0 c-vb ML/SM Emb Igneous. 19.8 95 4.8 2H to 1V 1.5H to 1V upstream and downstream slopes. Constructed SM/ML 35 to 55 < 10 3 to 8 unk unk probably comprised clayey alluvium with a
Colluvial moistened, then compacted by
1915 to 1920. high organic content. Upper portion from
horses and wagons.
colluvial/fluvial silty sands.

Sedimentary - 3H to 1V (8H 2 to 2.5H to 1V Zoned earth and rockfill dam with very broad core Well compacted, medium plasticity, sandy
Zoned 56 (slide 23 (8 to 36) - Rolled, well compacted.
USA, Emb & marine deposits. to 1V from 44 (6H to 1V from section (2H to 1V upstream and 1H to 1V Alluvial and CL (some SC) 66 (35 to clay with high undrained shear strength and
San Luis Dam 04-Sep-81 14 Earth and 4,2,2 c-vb CL area), 116 5650 48.7 unk medium 102.0% 1.2% dry Placed in 150 mm layers and -9
California Fndn Shales, sandstone m below 44 m below downstream) and thin outer earth and rockfill Colluvial - sandy clay 78) low permeability (10 m/sec). Placed dry of
Rockfill dam plasticity compacted by tamping rollers.
and conglomerate. crest) crest) zones. Constructed early to mid 1960's. OMC.

6 to 14 years
after dam Homogeneous earthfill embankment. Initially No formal compaction. 1896
Oct 1936, Sedimentary - 1896 fill poorly compacted. 1929/30 fill rolled
USA, raising. (40 to 0,0,0 or Emb & constructed in 1896 (to 13.7 m). Raised on CL - sandy 25 (medium fill compacted by stock.
Bear Gulch Nov 1942, Homog c-vb CL sandstone and 19.4 220 11.3 3H to 1V 2H to 1V Alluvial 65 20 to 25 unk unk (downstream raise). Sandy clay of medium
California 48 years after 1,1,1 Fndn downstream side in 1929/30 to 19.4 m. Rockfill clay plasticity) 1929/30 fill - rolled in thin
Nov 1944. shale. plasticity with some gravel.
initial drains in downstream slope, but no details. layers with sheepsfoot roller.
construction).
Puddle clay zone on upstream side of core-
wall, 8 to 11 wide at base. Silty clay (of
Concrete 42 (29 at 1.5H to 1V Concrete corewall rockfill embankment with zone CL (?) - silty 3 No formal compaction.
1927 and Emb & Sedimentary - 1.3 to 1.4H to Alluvial (puddle 16.2 kN/m medium plasticity) placed well wet of OMC.
Old Eildon Dam Australia 2 to 4 corewall, 10,0,0 - - failure 995 23.7 (including two of puddle clay on upstream side of corewall. clay, some CH high 70 to 80 unk wet OMC Wetted and then trafficked
April 1929 Fndn shale 1V clay) (dry) Of soft strength consistency (Su as low as 20
rockfill section) benches) Virtually completed by end of 1925. clays with horse and dray
kPa), stiff in upper 5 m. May have softened
significantly on wetting.
Appendix D Page D14

Table D1.2 Case studies of slope instability of embankment dams – slides in the upstream slope during drawdown (Sheet 2 of 5).
EMBANKMENT MATERIALS FOUNDATION MATERIAL PROPERTIES
Material 2 (Shoulder fill) General Comments on Design and Foundation (General) Foundation (If foundation Involved in Slide)
Name
Compaction
ASCS Density Moisture n Moisture
Source n % fines CF (%) PI (%) Compaction Comments Soil / Rock Description Description ASCS Class % fines CF (%) PI (%) Density Comments
Class Ratio Content Content

Homogeneous embankment constructed of CH/MH


Waghad - - - - - - - - - earthfill. Placed wet without any formal Soil Murum over weathered bedrock - - - - - - - -
compaction.

Failure occurred in homogeneous section


Dam FD2 - - - - - - - - - constructed on steep left abutment. Natural slope Rock Interbedded sandstones and shales. - - - - - - - -
is 7 to 13 degrees in upstream direction.

Alluvial - CL/CH clay "Adobe Clay", stiff to


Homogeneous embankment constructed of medium hard, friable, 2 to 12 m thickness.
Dam FD3 - - - - - - - - - Soil - - - - - - - -
to high plasticity sandy clays. Underlain by Alluvial sand and gravel, then
weathered shale.

Homogeneous embankment constructed of low to


Mondely Dam - - - - - - - - - unk unk - - - - - - - -
medium plasticity clays compacted wet of OMC.

Zoned earthfill embankment with thick central core


Limited information. Shoulder fill
Sampna Tank unk CL/CH > 50 > 50 unk unk 36% unk (1H to 1V slopes). Outer earthfill of high clay unk unk - - - - - - - -
described as not free draining.
content and low permeability.

Rolled, limited compaction. Placed Colluvial material derived from limited


GC - clayey
in 300 mm layers, compacted by transport of local sandstone hillslopes. Narrow puddle core of low plasticity sandy clay and Alluvial (sand, gravel and clay) of 14 m
Pilarcitos Dam colluvial sandy gravel, 23 < 10 11 94% unk Soil - - - - - - - -
carts and hauling wagons then rolled Fine to coarse gravel size. Layers shoulder fill of clayey sandy gravel. thickness overlying bedrock.
low plasticity
with 3 ton roller. sloping toward core.

Homogeneous earthfill embankment constructed of "Blue clay", not sure of origin. No


Charmes Dam - - - - - - - - - Soil - - - - - - - -
medium plasticity sandy silty clays. information on properties.

Homogeneous embankment constructed of silty Alluvial (?) - silty clay of 2 m thickness


Enista Dam - - - - - - - - - Soil - - - - - - - -
clays. overlying clayey gravels.

Alluvial. Suspect relatively sandy or


-6
Homogeneous embankment constructed of low gravelly given permeability of 10 m/sec.
Telish Dam - - - - - - - - - permeability materials. Suspect filling dominantly Soil Query over whether or not upper silt layer - - - - - - - -
clayey. was adequately removed prior to
construction.
Homogeneous embankment constructed of silty -7
- - - - - - - -
Drenovets Dam - - - - - - - - - Soil Alluvial (?). Permeability of 2 x 10 m/sec.
clays. Rockfill zone at downstream toe.

Homogeneous embankment constructed of sandy - - - - - - - -


Mechka Dam - - - - - - - - - unk unk
silty clays.

Homogeneous embankment constructed of sandy


Sushitsa Dam - - - - - - - - - unk unk - - - - - - - -
silty clays.

Limited information. Reference to being


Zoned earthfill embankment construction using a
Malko Sharkovo Cl (?) - sandy Likely wet Suspect rolled, well compacted dense, low permeability and having - - - - - - - -
unk unk unk unk unk sandy clay of low permeability for the upstream unk unk
Dam clay of OMC. (given period of construction) excess pore pressures post
shoulder.
constructions.

Homogeneous embankment constructed of medium Section indicates the foundation is clay.


Wassy Dam - - - - - - - - - Soil - - - - - - - -
plasticity silty clays. No further information than this.

Homogeneous embankment constructed of sandy


Mount Pisgah silts. Report that the work was not well done and Shallow thickness (up to 3 m) of clayey
- - - - - - - - - Soil - - - - - - - -
Dam was carried out with disregard of good dam sand overlying bedrock.
engineering practice.

Homogeneous embankment constructed of low Shallow thickness (3 m) of clay over


Forsythe Dam - - - - - - - - - Soil - - - - - - - -
plasticity sandy silts and silty sands. bedrock.

Zone 3 - weathered Highly plastic slightly sandy clay slopewash of


Rolled, reasonable to well shallow thickness. Some variability with low
sedimentary and CL/GC/SC - Zone 3 material comprises a mixture of CH (CL and GC
compaction of Zone 3 (300 mm Very broad core zone of medium plasticity sandy Colluvial slopewash and residual soils of Colluvial
alluvial earthfill, mixture of weathered rock and alluvium, in places) - dry (when plasticity and gravelly zones. Dry and hard when
San Luis Dam variable variable variable 99% 1.4% dry layers rolled with tamping roller). clays with thin outer rockfill zones on upstream Soil shallow depth overlying weathered slopewash on 83 55 40 to 55 unk embankment constructed, fissured. Softened
Zones 4 & 5 - materials reasonably well compacted. slightly sandy placed)
Zones 4 & 5 - up to 900 mm, limited slope. bedrock. hillslope. -9
crushed/blasted (Zone 3) Rockfill zones are loose placed. clay significantly on saturation. Low permeability (10
rolling -10
basalt rockfill. to 10 m/sec).

Limited information. 1936 investigation (in


Interbedded soft sandstone and clay shale.
Homogeneous embankment constructed of medium upstream slope) encountered a "wet mushy clay
1936 investigation encountered soft
plasticity sandy clays with no formal compaction. stratum" immediately underlain by a permeable
Bear Gulch - - - - - - - - - Rock "mushy" clay stratum below upstream - - - - - - -
Raise in 1929/30, on downstream side, using rolled zone. Hole remained dry until this permeable
shoulder (0.9 to 1.2 m below foundation
medium plasticity sandy clays. stratum was encountered. Water level rose to
level).
that of the reservoir.

Sedimentary - soft
3
shales and 14 kN/m Rockfill dumped at its angle of repose Zone of medium plasticity puddle clay placed Alluvium soils (clays overlying sandy
gravel to No compaction. Hand loaded onto Soil (in slide
Old Eildon Dam sandstones, and very low very low unk (dry unk (38 degrees). Shale prone weathering upstream of core-wall. Rockfill dumped at its angle gravels) of 6 to 20 m depth overlying Alluvial clays CL to CH (?) high high (?) unk unk unk Alluvial clay of "firm" to stiff strength consistency.
cobble size wagons then dumped. area)
Metamorphic - hard density) and breakdown, sandstone not so prone. of repose. bedrock.
slate
Appendix D Page D15

Table D1.2 Case studies of slope instability of embankment dams – slides in the upstream slope during drawdown (Sheet 3 of 5).
AT AND POST-FAILURE
Slope Slide Dimensions Post-Failure Deformation Behaviour
Name Slide Classification
Failure Slide Location State
(Hutchinson, 1988) Width, W Length, L Depth, D Height Slope Volume Velocity Velocity
Geometry W/L W/D Displacement (m) Comments
(m) (m) (m) (m) (deg) (cu.m.) (mm/day) Category

4.8 m crest settlement, 10


Rotational (possibly part Failure of upstream slope. Deep seated failure that (Moderate to Suspect moderate to rapid velocity of movement given relatively large distance
Waghad Type 2 (?) first time 330 (?) 90 11 29 14 200000 3.67 30.00 m displacement on lower -
translational) extended back to downstream edge of crest. Rapid) travelled. No indication of velity or time period of movement.
slope.

Compound - possibly two Reactivation of movement of existing slide monitored during 1994 drawdown.
Failure in upstream slope. Crest of back-scarp 6.5 m below
Dam FD2 translational planes on Type 2 reactivation 85 55 5.5 23 26.6 18000 1.55 15.45 0.423 m 27 slow Movement did not start until reservoir drawn down 5.6 m. Slide moved as
crest level.
which sliding occurred. intact mass.

Cracking at crest observed during 1928 drawdown. Slide occurred during


Failure in upstream slope. Crest of back-scarp 7.9 m below 1.7 m at head of scarp, 6
Dam FD3 Rotational Type 1 first time 100 23 5 12.1 26.6 7000 4.35 20.00 unk suspect rapid 1931 drawdown. Large and likely rapid movement at the slide toe where it ran
crest level. to 12 m at toe.
out onto the 2H to 1V upstream slope.

0.4 m on first drawdown Failure developed slowly. Longitudinal cracks observed on crest followed by
Failure in upstream slope. Extends back to near the
Mondely Dam Rotational Type 2 (?) first time unk 30 7 8.7 18.4 - - - (1981), 0.2 m on second 3 (up to 30 ?) very slow to slow development of the failure surface. Similar slow movements occurred on the
centreline of the crest.
drawdown. second drawdown.

1964 slide described as having "suddenly collapsed". Initial width 60 m,


Failures in upstream slope. Both 1961 and 1964 slides first time
Sampna Tank Rotational (?) Type 1 (?) 100 30 (g) 5 to 8 (guess) 15 26.6 10000 3.33 15.38 11.9 m - rapid spread to 90 m in a couple of hours. Suspect rate of movement was in the
extended back to the upstream edge of the crest. (1964 slide)
rapid category.

Failure in upstream slope on rapid drawdown. Top of back Upstream slope failure on very rapid drawdown (7.6 m in 2 weeks, 0.54
1 m at backscarp, only
Pilarcitos Dam Rotational (?) Type 1 (?) scarp about 3 to 4 m below crest. Several backscarps on first time unk unk unk unk 21.8 unk - - - suspect slow m/day). No information on velocity or timing of movement. Suspect slow given
rough estimate.
upper upstream slope. limited amount of deformation.

Failure in upstream slope. Extended back to the centreline


No information on amount and velocity of movement. Sherard (1954) indicates
Charmes Dam Rotational Type 2 of the crest. Possibly reasonably deep seated within first time unk 25 8 17 33.7 unk - - unk. - slow
movement occurred over a period of 6 weeks, therefore categorised as slow.
embankment, but this is not confirmed.

On large drawdown cracking was observed to have occurred over a width of


Failure in upstream slope. Top of backscarp 10 to 12 m 0.4 m (width of cracks on
Enista Dam Rotational (?) Type 2 (?) first time 200 60 unk 20 18.4 unk 3.33 - unk suspect slow 200 m on the mid to upper upstream slope. Only report on amount of
below crest on upstream slope. Possibly an incipient failure. upstream slope)
deformation is 400 mm width of cracks.

In 1977, on a routine drawdown, longitudinal cracks of 0.2 to 0.3 m width were


20 observed on the upstream slope, at 10 m below crest level. The cracks were
Failure in upstream slope. Top of backscarp 8 m below
Telish Dam Rotational (?) Type 2 (?) first time > 100 < 60 unk (maximum 18.4 unk > 1.7 - 1.0 m. unk suspect slow repaired. In 1980, on drawdown, wider and longer cracks were observed at 8
crest on upstream slope.
likely) m below crest (also repaired). In 1982, still wider cracks and slide movements
up to 1 m were observed.

Failure in upstream slope. No indication of location or size No information on slide. Suspect that it occurred during a relatively routine
Drenovets Dam Rotational (?) - first time unk unk unk unk 18.4 unk - - unk. unk suspect slow
of failure. drawdown.

Failure in upstream slope. No indication of location or size


Mechka Dam unk - first time unk unk unk unk 18.4 (?) unk - - unk. unk suspect slow No information on failure. Occurred during a routine drawdown.
of failure.

Failure in upstream slope. No indication of location or size


Sushitsa Dam unk - first time unk unk unk unk 18.4 (?) unk - - unk. unk suspect slow No information on failure. Occurred during a routine drawdown.
of failure.

Malko Sharkovo Failure in upstream slope. No indication of location or size


unk - first time unk unk unk unk unk unk - - unk. unk unk No information on amount of movement or velocity.
Dam of failure.

Failure of upstream slope on rapid drawdown (first


Wassy Dam Rotational Type 2 first time unk 25 6.5 16.5 33.7 - - - unk. unk unk No indication of velocity or amount of movement.
drawdown). Top of back-scarp at upstream edge of crest.

Mount Pisgah Failure of upstream slope following rapid drawdown. Top of 7 m at crest, up to 9 m at suspect No information on velocity of movement or comment on observations prior to or
Rotational Type 2 (?) first time 60 40 11 23.2 33.7 11000 1.50 5.45 unk
Dam back-scarp at downstream edge of crest. toe. moderate to rapid leading up to failure.

Limited information. Section showing post-failure profile is very approximate.


Failure of upstream slope following very rapid drawdown. 4 m at crest. Possibly 22
Forsythe Dam Rotational (?) Type 1 (?) first time unk 28 8 14.5 26.6 unk - - - suspect rapid Possible that the slide material that evacuated from the failure bowl moved
Top of back-scarp shown at centre of the crest. m at toe.
rapidly on the 2H to 1V slope accumulating at the upstream toe of the slope.

Slide first observed on 14 September 1981. Photos showed scarp evident on


Compound - translational
4 Sept. Suspect movement started much earlier (early August ?). Movement
basal component rotational Failure in upstream slope following rapid drawdown. Top of 15.8 (upper
San Luis Dam Type 2 first time 360 155 25 56 1.0E+06 2.32 14.40 17 m crest to toe region. ~ 1000 to 2000 Moderate likely to have initiated at a slow rate and gradually increased to a peak (about 1
backscarp through core back-scarp located 4.5 m below crest on upstream slope. slope 18.4)
to 2 m/day) over a period of 30 to 50 days before decreasing in rate and
material.
eventually coming to rest some 100 to 130 days after movement began.

Compound - translational Small slide in upstream slope in 1914 on very rapid drawdown (reservoir
Failures in upstream slope following drawdowns in 1936,
basal component in 1936 - 0.29 m; 1942 - emptied in 3 days). In 1936 a large crack was observed in upstream slope
1942 and 1944. Top of back-scarp located 3 to 6 m below
Bear Gulch foundation. Steep Type 2 first time 30 50 11.5 16.5 18.4 8000 0.60 2.61 0.32 m at crest unk suspect slow extending down the upstream slope to the toe. Similar movement occurred in
crest on upstream slope. Movement vector at top of
backscarp through backscarp. 1942, with the crack at a higher elevation. In 1944 - identical crack to 1942.
backscarp relatively flat (15 to 18 degrees).
embankment. Movement vector at backscarp 15 to 18 degrees to horizontal.

9.1 m crest settlement up


Compound - translational Deep seated failure in upstream slope on drawdown. Top of first time to Dec 1928. Apr to May Crest subsidence occurred prior to slide in 1929, and totaled 9.1 m. Suspect
220
basal component on clay backscarp is upstream side of corewall (centreline of crest). (1927), 1929 - 14 m crest triggered by 1927 drawdown. Movements, at a moderate velocity, occurred in
Old Eildon Dam Type 2 (widened to 50 26 29 36 300000 4.40 8.46 > 5000 (max) Moderate
(alluvium) foundation or Slide located over alluvial clay foundation between bedrock reactivated settlement, 17 m late April 1929 following drawdown and continued for several months at a
365 m)
layer immediately above. foundation area and lower hillslope of the left abutment. in 1929. displacement at upstream relatively slow rate.
toe.
Appendix D Page D16

Table D1.2 Case studies of slope instability of embankment dams – slides in the upstream slope during drawdown (Sheet 4 of 5).
POST FAILURE HYDROGEOLOGY
Degree of Reservoir Operation Pore Pressures
Name Break-up of
Comments Drawdown Causing Failure Comments
the Slide Steady /
Mass Comment Upstream Core Comments
Fluctuating
Standpipes indicate embankment is relatively permeable.
Limited information. Section indicates relatively large suspect limited response to reservoir (low limited change (low Query this given CH/MH material classification. Possible
Waghad minor to partial unk Maximum stored level at 24 m. Water level at end of drawdown at 10 m (possible 14 m drawdown) -
movement and partial degree of break up. permeability fill) permeability fill) that outer zones are permeable due to shrink/swell related
cracking.
1993 drawdown was largest and fastest (up to 3 times the rate
Typical yearly drawdown of 4 to 6 m Some response, but delayed compared to
First time slide suspected of occurring in 1993 (fastest previously) and suspect that slide occurred during this event. 1977 and
from full supply level. Significantly reservoir level. For 1977, 1988/89 and Small response to reservoir Permeable bedrock on downstream side of dam section.
Dam FD2 Virtually intact drawdown). Movement slowed once rate of drawdown Fluctuating 1989 events were to about the same base level as in 1993. -
larger drawdowns in 1976/77, 1993 drawdowns piezometric head in fluctuation. Zone 1 (or core) filling of low permeability.
decreased. 1994 - 10.6 m @ 270 to 300 mm/day; 1993 - 16.7 m @ 335 to 400
1988/89, 1992, 1993 and 1994. outer fill greater than reservoir level.
mm/day; 1976/77/88/89/92 - 9 to 14.5 m at 100 to 130 mm/day.

Typical yearly drawdown of 4 to 6 m, 1931 - 10.3 m at 110 mm/day (max) to base water level (19.4 m below Rainfall was not a contributing factor to instability
Significant, Slide material travelled down 2H to 1V slope onto flat berm Negligible change. Standpipe and Embankment is of very low permeability as indicated by
mostly from near full supply level. crest). Negligible response to (i.e. through water filled cracks). Limited rainfall for
Dam FD3 particularly at and then onto 5H to 1V slope. Note that different plans show Fluctuating piezometer records indicate negligible negligible change in piezometer and standpipe levels to
Large drawdowns (> 10 m) in 1928, 1928 - > 11 m at max 130 mm/day; 1930 - > 10 m at max 95 mm/day reservoir drawdown. 12 months prior to failure and no significant heavy
toe. differing amounts of movement at the toe. response to reservoir change. reservoir fluctuations.
1930 and 1931. (Note: max is average over 2 month period) rain days leading up to slide.

Drawdown failure in upstream slope that occurred very Very high pore pressures within embankment fill
Negligible change. Piezometers in -
slowly. Stabilisation works in early 1982 did not stop First drawdown in history of embankment. Occurred 1 year after the end Negligible response to Embankment materials of very low permeability (k = 2 x 10 developed during construction (ru ~ 0.8). Very slow
Mondely Dam Virtually intact Fluctuating no information upstream slope show little change to 11
additional movements occuring on the drawdown later that of construction. reservoir drawdown. m/sec). dissipation and still quite high at time of drawdown
fluctuations in reservoir level.
year. failure.

No information on 1961 drawdown.


1964 drawdown was the lowest drawdown in embankment Suspect 1964 - 9.1 m drawdown to 0.6 m above low water level (11.4 m below Suspect limited change given low Low permeability of outer zoned fill. High phreatic surface
Sampna Tank Virtually intact no information Suspect negligible. -
history. fluctuating crest, 9.6 m below maximum level) over 8 months at rate of 30 to 60 permeability of outer earthfill. in upstream shoulder on drawdown.
mm/day.

Expect shoulder fill to have relatively high permeability Suspect most rapid drawdown in embankment
given grading. Likely to have lower vertical permeability history. This drawdown was undertaken at the start
Reservoir drawn down very rapidly in 1969; 7.6 m in 2 weeks (540 Likely to follow reservoir level under
Pilarcitos Dam Virtually intact Limited information. unk no information Suspect limited. than horizontal due to layering during construction. Also, of the rainy season. May also have permeability
mm/day). Suspect most severe drawdown in history of dam. normal drawdown conditions.
sloping of layers toward core likely to result in perched restriction on upstream face due to siltation effects
water tables on very rapid drawdown. over the previous 100 + years.

Suspect limited response to reservoir


Suspect limited to Slide occurred at time of the deepest drawdown in
Charmes Dam unk Limited information. Likely slow velocity of movement. unk no information Deepest drawdown in history of embankment. change given embankment constructed of Likely low permeability of embankment fill.
negligible. embankment history.
medium plasticity sandy silty clay.

Limited reduction in phreatic surface to


Possibly an incipient failure. No indication that movements Lowest reservoir level prior to slide
Suspect Deepest drawdown in 13 year history of embankment. Reservoir drawn reservoir drawdown. Phreatic surface in Low permeability of embankment fill. Very limited response
Enista Dam Intact occurred as shearing on a defined surface of rupture (i.e., no occurring was to 11.2 m below crest Likely negligible. -
fluctuating down to 24.3 m below crest (22 m below maximum storage level). embankment only dropped 0.5 to 1 m from of phreatic surface in upstream shoulder to drawdown.
report on vertical displacement at back-scarp). (or 8.9 m below max. level).
level prior to drawdown.

1977 and 1980 cracking are indicative of an incipient failure Reservoir undergoes seasonal Negligible response to drawdown. 1983
1977 - routine drawdown. Suspect 1980 and 1982 drawdowns were also Low permeability of embankment fill. Negligible response
Telish Dam Intact (?) and were a precursor to the slide movements that occurred in Fluctuating drawdown. No indication of typical drawdown - over 3 to 4 months phreatic Likely negligible. -
relatively routine. of phreatic surface in upstream shoulder to drawdown.
1982. height or rate of drawdowns. surface only dropped 0.2 to 0.3 m.

Suspect it is much like the failure at Telish Dam. No Suspect Suspect limited to negligible response to
Drenovets Dam Intact (?) no information No quantitative information, but suspect the 1984 drawdown was routine. Likely negligible. -
information given on the slide. fluctuating reservoir fluctuations.
Low permeability of embankment fill, therefore likely limited
Suspect it is much like the slide at Telish Dam. No Suspect No quantitative information, but reference to 1981 drawdown as being Suspect limited to negligible response to
Mechka Dam Intact (?) no information Likely negligible. to negligible response of phreatic surface in upstream -
information given on the slide. fluctuating routine. reservoir fluctuations.
shoulder to drawdown.

Suspect it is much like the slide at Telish Dam. No Suspect No quantitative information, but reference to 1981 drawdown as being Suspect limited to negligible response to
Sushitsa Dam Intact (?) no information Likely negligible. -
information given on the slide. fluctuating routine. reservoir fluctuations.

Limited to negligible response to reservoir Low permeability of upstream shoulder fill as indicated by
Malko Sharkovo Failure of the upstream slope occurred during the first Failure on first
unk Failure on first drawdown Failure on first drawdown. No quantitative information on the drawdown. fluctuations as indicated by slow Likely negligible. presence of construction pore pressures 1 year after -
Dam drawdown. Construction pore pressures still present. drawdown
dissipation of construction pore pressures. construction had been completed.

Was the first drawdown and was undertaken within a year of completion Little time for infiltration of stored water No opportunity for a Likely low permeability of embankment filling. If placed wet Failure occurred at time of very rapid drawdown.
Failure on first
Wassy Dam unk Slide occurred on first drawdown, which was very rapid. Failure on first drawdown of construction. Very rapid drawdown of 10 m within 20 to 25 days (400 into the medium plasticity sandy clay response given short of optimum may have high excess pore pressures Reservoir had only reached full supply level in early
drawdown
to 500 mm/day) from full supply level. embankment fill. impoundment period. associated with construction. August 1883 before it was drawn down.
Likely that embankment fill is of moderate permeability and
Reservoir full in early summer months (June to July). Drawdown of 13.4
Slide occurred following a rapid drawdown of the reservoir. Some but likely limited reduction in would respond with reservoir under slow drawdown Failure occurred following very rapid drawdown over
Mount Pisgah m over about two months initially at a steady rate and then at a very rapid Suspect some response but
Some break up Suspect of moderate to rapid velocity given the amount of unk no information phreatic surface to reservoir drawdown, conditions (contains 70% coarse silt to sand size and low a period of several days on the back of a steady
Dam rate for several days prior to the failure. Average rate for drawdown is likely delayed.
movement that occurred. particularly at high drawdown rates. clay content). Suspect delayed response under rapid drawdown over several months.
220 mm/day.
drawdown.
Failure triggered by extremely rapid, but short, drawdown.
Possibly slide occurred Erosion through spillway occurred Once reservoir level reached full supply level erosion in the spillway Negligible response likely
Suspect failure confined to the mid to upper portion of the Negligible response likely under the No time for drainage to occur in the short time (2 hours) it Reservoir impounded some water during
Forsythe Dam significant break on first shortly after reaching full supply region washed out a 4.6 m deep channel. The reservoir level was drawn under the extremely rapid
upstream slope and the run-out from the failure travelled extremely rapid drawdown. took for the reservoir level to fall 4.6 m. construction.
up. drawdown level. down 4.6 m in several hours; i.e., at an extremely rapid rate (2.3 m/hour). drawdown.
down the steep upstream slope.

The slopewash (or portions of it) and the upstream portion


Typically, seasonal drawdown of 10 Zone 3 - responds with reservoir (some
Monitoring only picked up the latter part of movement, the of the Zone 1 fill show a delayed and generally some to
to 25 m (years preceeding slide). 1981 - Drawdown from full supply level to EL 110.5 m (55 m) in 119 days piezos show a delay). Zone 1 - upstream Zone 1 - delayed, very Failure occurred following the most rapid and largest
slow down phase after the peak velocity. Peak velocity likely limited response to reservoir fluctuations. On large and
Larger events in 1975 (30 m at 350 at a maximum average 50 day rate of 625 mm/day. Largest drawdown in portion shows some to limited and delayed limited to negligible drawdown in the history of the embankment. The
San Luis Dam Virtually intact on or slightly before the 14 Sept 1981 (some 30 to 50 days Fluctuating rapid drawdowns therefore, the core and slopewash show a
mm/day), 1976 (43 m at 560 embankment history. Only 1977 drawdown was to a lower base water response. response to reservoir reservoir had been subjected to several previous
after movement likely to have started). Movements show piezometric level above that of the reservoir. (Note that
mm/day) and 1977 (36.5 m at 450 level (EL 107.5 m). Slopewash - variable response, some fluctuations. large and rapid drawdowns.
similar velocity on different parts of the slide surface. reference to the slopewash is in the vicinity of the slide
mm/day). piezos show lag with some response.
area).

Small amount of movement would suggest slow velocity of


1936 - 8.2 m drawdown at average of 90 mm/day, but occurred early in Unknown, suspect limited response given
movement. Small amount of "mushy" material material in No information on piezometric levels. Suspect some but
suspect virtually Typical drawdown rate of about 90 the season. fill is medium plasticity sandy clay. Suspect limited to
Bear Gulch crack (backscarp), but otherwise "sound and dense" material Fluctuating delayed response to drawdown given embankment fill is a -
intact. mm/day. 1942 - similar drawdown rate and to similar level as in 1936. Permeable zone in foundation follows negligible.
either side of crack. Movement suggests strongly medium plasticity sandy clay.
1944 drawdown similar again. reservoir.
translational with formation of tension crack at backscarp.
Suspect large drawdown in 1927 initiated the slide, and that
Reservoir in operation whilst
much of the crest subsidence prior to 1929 occurred during Suspect initial failure occurred following the very
embankment under construction. 1927 - 32.5 m at average rate of 190 mm/day. Rate up to 500 mm/day in Puddle Clay - limited
this event (reference to continued dumping of rockfill at crest Rockfill permeable. Pore pressures from records covering 1944 to 1948. large and rapid drawdown of 1927. Significant
Water impounded from 1922 and latter weeks of drawdown. response to reservoir level
Old Eildon Dam Minor break up to maintain elevation). In 1929 the movement was Fluctuating Foundation - variable response, typically Elevated pore pressures in foundation and puddle core on movement occurred on the much smaller 1929
subject to seasonal drawdown. 1929 - 15.5 m at average rate of 100 mm/day. Increased rate of 230 fluctuations. Trend follows
reactivated on drawdown. Widening of the crest (post the delayed response of lesser magnitude. large and rapid drawdowns. drawdown. In between drawdowns the crest of the
Reached full supply level in August mm/day in 2 weeks prior to slide. reservoir level but delayed.
1927 drawdown) attributed to decreasing the overall stability upstream slope was significantly widened.
1927.
of the slope.
Appendix D Page D17

Table D1.2 Case studies of slope instability of embankment dams – slides in the upstream slope during drawdown (Sheet 5 of 5).
MECHANICS OF FAILURE
Name Other Comments References
Cause/s Landsliding Trigger Failure Mechanism Comments on Failure Mechanics

Penman (1986)
Initiated as drained failure. Large movement of relatively flat upstream slope suggests Embankment materials are poorly compacted CH/MH placed well wet of OMC. Suspect that the earthfill is strain No indication of previous movement during earlier
Waghad Drawdown Drawdown Nagarkar et al (1978)
embankment materials are significantly strain weakening. weakening on shearing from fully softened to residual strength. drawdowns. Only limited information available.
Nagarkar et al (1981)

Piezometers in outer zone of Zone 1 filling (core material in zoned section and material used in homogeneous section)
Drawdown induced failure occurred on left abutment
Drawdown, large and relatively rapid. Initiated in drained loading. Possible progressive undrained weakening under high shear stress indicated slow response to reservoir drawdown. Suspect instability associated with high piezometric levels in earthfill on
Drawdown, 1993 likely event (constructed on upstream sloping (7 to 13 degree slopes)
Dam FD2 1993 event largest drawdown and conditions from large drawdowns prior to 1993. Initial slide may have occurred in 1993 following large drawdown. Zone 1 material is dilative under low confining pressures and undrained strain weakening of dilative
during which failure occurred. hillside). Embankment design on left abutment is
fastest rate (~ 3 times previous rates). progressive weakening from 1977 and 1988/89 drawdowns. earthfill may have occurred under low confining pressures. Mechanism could incorporate progressive failure due to strain
homogeneous earthfill at 2H to 1V slope.
weakening on previous large drawdowns.

1931 Drawdown, large and rapid. Embankment fill is of very low permeability and shows little response to reservoir drawdown. Under drained conditions the Rapid failure of upstream slope on drawdown. Cracking in
Slide initiated as drained failure. Potential for progressive undrained strain weakening of fill
Relatively steep upstream slope (2H to fill would be dilative and strain weakening at low confining pressures. Previous large drawdowns possibly resulted in 1928 indicative of marginal stability of slope to rapid and large
Dam FD3 Drawdown of 1931. under high shear stress conditions from previous of large and rapid drawdowns of 1928 and
1V) for embankment constructed of strain weakening, as indicated by cracking that occurred following 1928 slide. Restraint from concrete facing and piling drawdown. More recent large drawdowns have resulted in
1930. Concrete facing and piling contributed to brittleness of slide.
medium plasticity clays. contributed to brittleness of slide mechanism and rapid post failure velocity of the slide. cracking of the embankment.

This failure may be similar in some respects to that of Lake


Drawdown. High construction pore Suspect undrained failure through homogeneous fill triggered by drawdown. Suspect that Alonso et al (1997) comment that they obtained a factor of safety of close to 1 for drained analysis (using pore pressure
Shelbyville. Similarites are timing, velocity of movement and Alonso et al (1997)
Mondely Dam pressures in fill (placed wet of OMC First drawdown. embankment in limiting stability condition at the end of construction. Construction pore pressures recorded from piezometers) but greater than one for undrained analysis. Their analysis would indicate the embankment
involvement of first drawdown. Very slow to slow failure in Lino (1997)
and over-compacted). were a significicant factor in the failure. was in a limiting stability condition at the end of construction.
upstream slope on first drawdown.
Relatively steep upstream slope would have contributed to rapid rate of movement as material evacuating failure bowl
Initiated as a drained failure. Large and rapid movement suggests embankment materials are Rapid failure of upstream slope on drawdown. Possible
Drawdown, low permeability of would break up and move relatively quickly on the 2H to 1V upstream slope, therefore having continued loss of toe
significantly strain weakening on shearing. Brittleness possibly associated with surface of progressive strain weakening from previous drawdowns.
Sampna Tank upstream shoulder fill (high phreatic Drawdown support. Dominant influences on failure is the low permeability of shoulder fill and steep upstream slope. Suspect filling (if Sinha (1968)
rupture passing through partially saturated soils. Possible progressive weakening from earlier Slope too steep for low permeability fill used in upstream
surface). reasonably well compacted) would be dilative and strain weakening in drained and undrained loading at low to moderate
drawdowns. shoulder.
levels of confining pressure.
Very rapid drawdown. Possible
perched water tables due to layering Drained failure in dilative outer shell of embankment (clayey sandy gravel). Suspect pore Upstream shoulder fill would be dilative on shearing at low confining pressures. Likely that this dilative behaviour limited Failure in upstream slope of clayey sandy gravel during very
Sherard (1953)
Pilarcitos Dam sloping toward core. Rainfall during Very rapid drawdown. pressure within shoulder could not drain at same rate as reservoir leading therefore to slope the amount of shear movement. Downstream sloping layering toward core likely to have restricted drainage and rapid drawdown. Limited details, but likely low rate of
USCOLD (1975)
drawdown may also have had an failure. maintained phreatic surface at relatively high levels during very rapid drawdown. movement, possibly in slow category.
effect.
Drawdown, low permeability of Suspect fill to have limited dilatancy on shearing at low confining pressures (due to likely limited strain weakening).
Slow failure on deepest drawdown in embankment history in
upstream shoulder fill therefore likely Initiated as a drained failure. Slow movements suggest embankment materials are not Possible progressive undrained weakening occurred as a result of high shear stress levels associated with previous
Charmes Dam Drawdown steep upstream slope (1.5H to 1V) of medium plasticity sandy Sherard (1953)
high phreatic surface retained in significantly strain weakening on shearing. drawdowns. Would expect high shear stress to be present in upstream slope on drawdown due to steepness of upstream
silty clay.
embankment. slope (1.5H to 1V).
Drawdown. Low permeability of fill Further cracking occurred under drawdowns after 1982 (when stabilising berms had been added). This may indicate that
Drained instability in upstream slope on large drawdown due to retention of high phreatic surface Possible incipient failure that occurred during a large
resulting in limited reduction in phreatic Drawdown (largest by some 13 m the failure in 1982 was in the initial stages of failure and the dilative tendancy of the fill on shearing and its low permeability
Enista Dam in upstream shoulder. Likely dilation on shearing resulted in limited and likely slow velocity of drawdown (by far the biggest event in the embankments Abidjiev (1994)
surface in upstream shoulder of the in embankment history) limited the amount of deformation. The later movements may have occurred at a lower undrained strength due to strain
movement. history). Limited movement on a upstream slope of 3H to 1V.
embankment. weakening and dilation after the 1982 event.

Drawdown. Low permeability of fill Initiated as a drained instability, but development of negative pore pressures on shearing in Drawdown failure in 1982 under routine drawdown operation.
resulting in negligible reduction in the Drawdown (likely that was only a dilative fill may have limited the movement. Further movements under later drawdowns suggest Cracking in 1977 and again in 1980 were an indication of the progressive strain weakening of the embankment fill. The Preceeded by cracking in 1977 and 1980, also under routine
Telish Dam Abidjiev (1994)
phreatic surface within the upstream routine drawdown) progressive strain weakening of fill under the high shear stress conditions associated with filling is likely to be dilative and strain weakening at low confining pressures. drawdowns. Failure indicative of progressive undrained strain
shoulder of the embankment. seasonal drawdown. weakening under routine (but large) operational drawdowns.

Limited information. Failure in upstream slope (3H to 1V),


Drenovets Dam Drawdown, possibly routine.
possibly during routine drawdown.
Drawdown. Low permeability of fill
Suspect that given the similarities in the period of construction, embankment design, material
resulting in limited reduction in phreatic Limited information. Failure in upstream slope during routine
Mechka Dam Drawdown, possibly routine. type, embankment height, upstream slope and seasonal operation of the reservoirs, that the - Abidjiev (1994)
surface in upstream shoulder of the drawdown.
failures of Telish, Drenovets, Mechka and Sushitsa are all of similar mechanism.
embankment.
Limited information. Failure in upstream slope during routine
Sushitsa Dam Drawdown, possibly routine.
drawdown.

Very limited information. Upstream failure during first


Drawdown. Presence of residual Pore pressures developed during construction indicate placement on the wet side of optimum and therefore potentially drawdown in zoned earthfill embankment with upstream
Malko Sharkovo Possible undrained failure through cohesive upstream shoulder fill triggered by drawdown. Likely
excess pore pressures from First drawdown. relatively low undrained strength (depending on placement water content). It is possible that the embankment was in a shoulder of low permeability sandy clay. Construction pore Abidjiev (1994)
Dam that the embankment was in a limiting stability condition at the end of construction.
construction in upstream shoulder fill. limiting stability condition at the end of construction. pressures were still present in the upstream slope at the time
of failure (1 year after construction).
If an undrained failure, would require placement of filling wet of optimum and limiting stability condition at the end of Interesting failure. Suspect either undrained failure of wet
Drawdown, very rapid and first.
Drawdown, first drawdown and Several possibilities. Possible undrained failure triggered by drawdown. Possibly initiated as a construction (similar to Mondely). Another possibility is that filling was poorly placed (with no moisture control and/or in compacted filling of low undrained shear strength or drained
Wassy Dam Possibility of residual pore pressures Sherard (1953)
very rapid. drained failure if filling poorly placed and softened on wetting. thick layers) resulting in relatively high horizontal permeability so that on reservoir filling water could readily penetrate the failure of poorly compacted filling that softened significantly on
associated with construction.
filling and cause softening. wetting.
Suspect initiated as a drained failure. Large movement suggests likely rapid velocity of the slide
Rapid drawdown. Delayed response The steep upstream slope is likely to have had a significant influence on the likely rapid velocity of movement. Once the Likely rapid failure of upstream slope on very rapid drawdown.
Mount Pisgah mass, possibly indicating significant brittleness of the slide mass. Contributing factors to the Sherard et al (1963)
of phreatic surface on very rapid Rapid drawdown. slide had initiated and the concrete facing was no longer supporting the steep upstream face, the slide occurred relatively Embankment constructed of sandy silt with steep upstream
Dam brittleness would be the reinforced concrete upstream facing. The steep upstream slope is also a Sherard (1953)
drawdown. quickly and travelled a large distance. face (1.5H to 1V). Poor quality of workmanship.
contributing factor to the large movements.

Limited information. Likely rapid failure of mid to upper


Initiated as a drained failure, under extremely rapid drawdown conditions, of relatively steep (2H Possible that failure may have occurred as a series of thin sloughing type failures to give post-failure profile of slide shown
Forsythe Dam Extremely rapid drawdown. Drawdown, extremely rapid. portion of the upstream slope on an extremely rapid Sherard (1953)
to 1V) upstream slope. by Sherard (1954).
drawdown.

Initiated as a drained failure and progressed under undrained conditions. It is likely that The core material was rolled and well-compacted and would therefore have a high undrained shear strength at
progressive strain weakening of the core and slopewash occurred under the 1981 and previous compaction. At relatively low confining pressures softening would occur due to swelling and saturation, however, it would
Well documented and analysed case study of failure during
Drawdown of 1981. Elevated large drawdowns under the high shear stress conditions imposed under drawdown. The still be dilative on shearing under these conditions. Further strain weakening (in undrained loading) potentially occurred
drawdown. Failure occurred on the largest and most rapid Stark and Duncan (1987,
piezometric levels in slopewash and Drawdown of 1981, quickest and slopewash also softened significantly on saturation, and in the failure area is likely to have under high stress conditions. Suspect that at failure that the strength of the core material was somewhere between peak
San Luis Dam drawdown in the embankments history. Progressive 1991)
core. Strain weakening of both largest in embankment history. resulted in shear straining under the stress levels imposed. Further shear straining is likely to and fully softened. For the slopewash it is possible that the strength at failure was somewhere between fully softened and
weakening from previous large drawdowns could have been a Von Thun (1988)
slopewash and core. have occurred under the higher stress conditions imposed from previous large drawdowns. residual (depending on the amount of shearing and degree of localisation). Restraint on the lateral margins is a
significant factor in the mechanics of failure.
Likely significant lateral restraint on margins. Favourable orientation of the natural hillslope also significant factor. The slide was hinged about the southern end and the width of the slide gradually increased over time in
a contributing factor. the southern direction. Changes in the natural topography contributed to the restraint on the southern lateral margin.

Suspect possible progressive weakening within a layer in the foundation under drawdown.
Movement vector at backscarp suggests strongly translational movement of slide. Movements Surface of rupture is not confirmed through foundation, but is
Bear Gulch Drawdown. Drawdown. - Sherard (1953)
occurred when the reservoir level was quite low suggesting possible high uplift pressures in toe likely to have been located on wet, soft clay seam.
region.

The chain of events are difficult to follow in the paper by Knight (1938) and the monitoring records indicate the crest settled
Slide in upstream slope on drawdown. Initiated in 1927 and
Rapid drawdown. Likely initiation Several possibilities. In 1927 it is possible that construction pore pressures were still present in up to 14 m (or 50% of its height) in about 20 to 25 days. For an additional 6 m of movement to occur following an initial 8 m
Rapid drawdown (1927 and 1929). reactivated on the 1929 drawdown. Impoundment during
following 1927 drawdown and the puddle core and possibly the foundation. Therefore, the failure may have initiated as an in late April 1929 would suggest that rockfill was being dumped in the crest region of the failure, effectively reducing the Knight (1938)
Old Eildon Dam Construction (widening of upstream construction is considered to have had an effect on stability
reactivation following 1929 undrained failure triggered by the drawdown in 1927. The movement in 1929 was a reactivation stability of the slide. Restraint on the lateral margins is evidenced by the widening of the slide from 215 to 365 m. Speedie (1948)
side of crest). under drawdown (possible that failure may have occurred
drawdown. of the existing slide. Significant restraint on the lateral margins. Therefore, spreading of load onto the margins of the slide occurred extending the width of the slide, as opposed to
during construction if water not impounded).
shearing through rockfill.
Appendix D Page D18

Table D1.3: Case studies of slope instability of embankment dams – post-construction slides in the downstream slope.
Embankment Zoning and MATERIALS
Years Since Slope Embankment Dimensions
Name Location
Date of
End of Classification Slide
Failure Geology Material 1 (Core)
Failure Location
Construction n Core Core Geometry Height, Length, Ratio Upstream Downstream ASCS % CF Density Moisture Compaction
Zoning Class Comments Source n PI (%) Compaction Comments
Width Type He (m) L e (m) Le/He Slope Slope Class fines (%) Ratio Content Rating
Glacial - moraine Homogeneous earthfill SC - silty
Park Reservoir USA,
May-69 60 Homog 2,0,0 c-vb SC Emb Type 1 and lacustrine 24.4 366 15.0 3H to 1V 1.5H to 1V embankment with rockfill toe, Glacial (?) and clayey unk unk unk unk unk unk unk No information.
Dam Wyoming
deposits. completed in 1909. sand.

Sedimentary - Homogeneous earthfill Well compacted sandy gravel with low


2.5 to 5H to Alluvial - 10 non Compacted with Rolled, well
Arroyito Dam Argentina Jan-84 6 Homog 1,0,2 c-vb GP/GW Emb Type 2 sandstones and 20 3400 170.0 2.5H to 1V embankment constructed 1974 to GP/GW unk DR = 100% unk content fines (maximum 10 % fines).
1V sandy gravels (max) plastic heavy rollers compacted
claystones. 1978. Stratification in embankment with kh > kv.

Probable thin
Homogeneous earthfill High plasticity silty clay. Potentially poor
moistened layers,
embankment, constructed in 1910. CH - silty No formal compaction if hard clays not broken up
Barton Dam USA, Idaho Jun-22 12 Homog 0,0,0 c-vb CH Emb Type 1/2 (?) Lacustrine 12.2 300 24.6 2.5H to 1V 1.5H to 1V Glacial 100 60 38 unk unk and compacted by
Relatively steep downstream clays compaction adequately prior to compaction. Possible
travel of teams and
slope, no filters. stratification (kh > kv) in embankment.
wagons.

Emb Puddle core earthfill embankment Refered to as well High plasticity clay of soft to firm strength
Harrogate Puddle core Glacial - Boulder Glacial - high No formal
UK 18-Dec-51 81 8,0,0 - CH (possibly Type 2 8.8 unk - 1.9H to 1V 1.9H to 1V with narrow core and no filters. CH high unk unk unk compacted for consistency (Su ~ 28 kPa). Narrow, steep
Dam earthfill clay Boulder Clay (?) compaction
Fndn also) Constructed 1870. period. sided central core zone.

No information on earthfill. Suspect clayey


Emb and of relatively low permeability given
Woodrat Knot USA, Homogeneous earthfill
May-61 5 Homog 0,0,0 (?) c-vb unk (possibly unk Tuff 26 230 8.8 2.5H to 1V 1.75H to 1V unk unk unk unk unk unk unk unk unk slow time for dampness to be observed on
Dam Oregon embankment, completed in 1956.
Fndn also) the downstream slope. Likely stratification
with kh > k v.

Placed in layers
Homogeneous earthfill dam.
2 (after raise to and compacted Medium plasticity silty clay. No formal
Emb Initial construction in 1898 (to 3 m) 17 -
Fruitgrowers USA, 11 m), 39 (after Sedimentary - Residual - 30 to with teams. No formal compaction and possibly placed dry of
12-Jun-37 Homog 0,0,0 c-vb CL (possibly Type 1/2 (?) 11 280 25.5 3H to 1V 2H to 1V then raised in 1905 (to 6.7 m), CL 98 medium unk unk
Dam Colorado initial shale soft shale 40 Reference to compaction OMC. For 1935 raise reference to
Fndn also) 1910 (to 9.8 m) and 1935 (to 11 plasticity.
construction) placement in dry moistening of fill.
m). All raises on upstream slope.
condition.

Likely limited
Homogeneous earthfill
Santa Ana formal High plasticity silty clays. Limited
Emb & embankment. Initial embankment CH - silty unk (high
Acaxochitlan Mexico Dec-52 21 Homog 0,0,0 c-vb CH Type 2 unk 12 unk - 2.5H to 1V 2H to 1V unk 85 50 unk unk unk compaction information. Not sure of borrow source,
Fndn failed in 1925. Completely clays plasticity)
Dam given period of possibly alluvial.
reconstructed, completed 1931.
construction.

24 to 38 (in Limited information. Swelling of


Homogeneous embankment (in residual - 3 embankment post construction an
21 - 2.03 t/m slide area),
Emb & Sedimentary - failed section) with foundation filter weathered Rolled, well indication of high compaction. Medium
Siburua Dam Venezuela 15-Jul-64 7 Homog 1,0,1 c-vb CL Type 5 14.4 900 62.5 2.5H to 1V 2H to 1V CL unk unk medium (bulk 21 to 25% unk
Fndn marine shales. under downstream shoulder. shales and compacted -
plasticity density) (outside slide plasticity shaly clays. Low permeability, 10
Constructed 1956 to 1957. shaly clay 10
area) m/sec.

Zoned earth and rockfill > 90%


High plasticity clay of stiff strength
1949 to embankment with broad clay core not known SMDD Suspect rolled,
Seven Sisters Zoned Earth Emb & 3 5 to 10% wet consistency (Suv = 50 to 145 kPa, avg =
Canada 1956 (13 0.3 to 7 4,0,1 c-vb CH Type 5 unk 7.3 5600 767.1 3.5H to 1V 2.5H to 1V (1.5H to 1V downstream and 2H to (alluvial or CH unk unk 35 (1.97 t/m unk but relatively
Dike and Rockfill Fndn OMC 100 kPa). Variable plasticity (liquid limit
failures) 1V upstream) and thin outer lacustrine) bulk poor.
range from 19 to 114).
rockfill zones. density)

Homogeneous earthfill
1 (after dam
embankment. No information on
USA, raise), 51 Emb & cross section geometry.
Great Western 15-Jun-58 Homog 0,0,0 (?) c-vb CL (?) Type 2 unk 18.6 580 31.2 unk unk alluvial (?) CL (?) unk unk unk unk unk unk unk Referred to only as rolled earthfill.
Colorado (since original Fndn
Combined upstream and
construction) downstream slope = 4.5H to 1V.

Zoned earthfill embankment with Suspect rolled,


Sedimentary - broad clay core (1H to 1V well Limited information. Suspect reasonable
alluvial - 1.44 t/m3
Zoned Emb & conglomerate and 2 to 3H to upstream and downstream). very compacted well compacted given the age at
Aran Dam India Apr-78 1 3,0,0 c-tk CH Type 2/5 (?) 30.3 750 24.8 2 to 2.5H to 1V black cotton CH high high (dry unk Rolled earthfill
Earthfill Fndn other sedimentary 1V Central cut-off through to below high (given period construction (1975 to 1977). High
soils density)
rocks. the conglomerate layer. of plasticity.
Constructed 1975 to 1977. construction)

Placed in thin
Homogeneous earthfill layers with scraper Constructed of slightly gravelly sandy clays
Lake Yosemite USA, Emb & Alluvial / CL/SC - 11 - low No formal
1943 60 Homog 0,0,0 c-vb CL/SC Type 2 unk 16.2 1500 92.6 2H to 1V 2H to 1V embankment constructed in 1883 50 10 unk unk teams and and clayey sands of low plasticity. Limited
Dam California Fndn (?) Colluvial gravelly plasticity compaction
to 1884. compacted by compaction.
travel of the teams.

Homogeneous earthfill Suspect no


Constructed of medium plasticity sandy
2 (after raise of embankment. Originally formal
15 - clays with limited compaction (for original
USA, 0.9 m), 41 Emb & constructed in 1910 to 6.7 m. In Colluvial and CL - sandy compaction
Yuba Dam Jan-51 Homog 0,0,0 c-vb CL Type 2 unk 7.6 270 35.5 2.5H to 1V 1.75H to 1V 78 25 medium unk unk unk construction). Downstream raised section
California (after initial Fndn 1949 raised (on downstream side) Residual clay (given period
plasticity of better compaction and lower
construction) by 0.9 m height. Relatively steep of
permeability.
downstream slope (1.75H to 1V). construction)
Appendix D Page D19

Table D1.3: Case studies of slope instability of embankment dams – post-construction slides in the downstream slope (Sheet 2 of 4).
MATERIALS FOUNDATION MATERIAL PROPERTIES
Name Material 2 (Shoulder fill) Foundation (General) Foundation (If foundation Involved in Slide)
General Comments on Design and
ASCS % CF Density Moisture Compaction Compaction Soil / ASCS Moisture
Source PI (%) Compaction Comments Description Description % fines CF (%) PI (%) Density Comments
Classn fines (%) Ratio Content Rating Rock Classn Content
Homogeneous embankment constructed of
Park Reservoir Glacial moraine or lucustrine
- - - - - - - - - - silty and clayey sands. No information on Soil - - - - - - - -
Dam deposits.
construction methods.

Homogeneous embankment constructed of


well compacted sandy gravels of low fines Sandy gravels to 10 m depth
Arroyito Dam - - - - - - - - - - Soil - - - - - - - -
content (maximum 10%). Stratification in overlying bedrock.
embankment.

Lacustrine clays, high plasticity,


Homogeneous embankment constructed of heavily over-consolidated and
Barton Dam - - - - - - - - - - Soil - - - - - - - -
CH lacustrine clays. very stiff to hard strength
consistency.

Glacial (?) - 22 to 25 19 kN/m3 22 to 38% Medium plasticity clay of soft to firm Narrow puddle core of CH clay and shoulder Deep clay profile. Surface soil
No formal
Harrogate Dam Boulder Clay CL unk unk (medium (bulk (post unk strength consistency (Su ~ 18 to 35 zones of medium plasticity clays. No formal Soil not removed. Clay of stiff to very - - - - - - - -
compaction.
(?) plasticity) density) failure) kPa). No formal compaction. compaction. stiff strength consistency (?).

Woodrat Knot Homogeneous earthfill embankment. Very


- - - - - - - - - - unk unk - - - - - - - -
Dam limited data. Suspect clayey earthfill.

Homogeneous earthfill embankment of


Fruitgrowers Alluvial (?) - clayey sand. No
- - - - - - - - - - poorly compacted, medium plasticity silty Soil - - - - - - - -
Dam details other than this.
clays.

Alluvial (?). No cleaning of


22 to 32 No foundation clean up.
Santa Ana foundation. Layer of high unk (likely
Homogeneous earthfill embankment Organic clay (liquid limit Embankment founded on a layer of
Acaxochitlan - - - - - - - - - - Soil plasticity organic clay (0.5 m OH 80 high unk that
constructed of high plasticity silty clays. layer = 70 to high plassticity organic clay
Dam thick) overlying "soft" clay. saturated)
114) underlain by a "soft" clay.
Below this, "compact" clayey silt.

Medium plasticity shaly clays.


Residual Soil - Shaly clay of 3 to
23 to 25% Permeable layers through the
Homogeneous embankment constructed of 4 m thickness overlying
21 - inside slip foundation, more so in the
medium plasticity shaly clays. In slide area, weathered shale. Bedding in
Siburua Dam - - - - - - - - - - Soil Shaly Clay CL unk unk medium unk area, weathered zone below the residual
foundation slopes in downstream direction at shaledips in N to NW direction at
plasticity ~ 21% soil. Likely relict bedding in residual
7 to 11 degrees. 17 degrees (favourable oriented
outside. soil profile oriented favourably in
to slide location).
slide location.

High plasticity, firm to stiff (Su = 28


rockfill - Alluvial (?) - high plasticity clay of
Zoned earth and rockfill embankment with 16.3 to 180 kPa, avg 88 kPa) clay.
Seven Sisters gravel to No information, referred to only as 3 to 6 m thickness underlain by High plasticity 3
unk unk unk unk unk unk unk unk broad CH clay core and thin outer rockfill Soil CH unk unk 59 kN/m bulk up to 45% Upper 1 m cracked and fissured.
Dike boulder size rockfill. low plasticity clay and/or clay
shells. unit weight Fissuring continues to 2.1 m.
(?) bedrock.
Stratified.

Homogeneous earthfill embankment. Very Alluvial - clay. No further


Great Western - - - - - - - - - - Soil - - - - - - - -
limited data. Suspect clayey earthfill. information than this.

Limited information. Suspect


reasonable well compacted given the
Alluvial and Suspect rolled, age at construction (1975 to 1977). Alluvial - fissured high plasticity High plasticity, highly fissured clay
1.76 t/m 3 1.33 t/m 3
Residual suspect GC well compacted High gravel content and likely low Zoned earthfill embankment with broad CH clay of 5 m thickness overlying Fissured, high foundation. Reference to having
Aran Dam unk unk unk (dry unk Rolled Soil CH high (?) high (?) high (?) (dry unk
(murum and to GM (given period of fines content. Phreatic surface clay core and gravelly earthfill shoulders. thin murum layer then plasticity clay. alow undrained shear strength when
density) density)
conglomerate) construction) shown through the downstream conglomerate bedrock. saturated.
shoulder would indicate this material
has a relatively low permeability.

Alluvial (?) - referred to as


Homogeneous earthfill embankment cemented gravels, but similar to
Lake Yosemite
- - - - - - - - - - constructed of slightly gravelly sandy clays Soil material of which embankment - - - - - - - Limited information.
Dam
and clayey sands. Limited compaction. constructed (slightly gravelly
sandy clays and clayey sands.

Homogeneous earthfill embankment


constructed of medium plasticity sandy clays
Alluvial (?) - clay, no further
Yuba Dam - - - - - - - - - - of likely no formal compaction. Raised zone Soil - - - - - - - Very limited information.
information.
on downstream side) of improved
construction practice, but lower permeability.
Appendix D Page D20

Table D1.3: Case studies of slope instability of embankment dams – post-construction slides in the downstream slope (Sheet 3 of 4).
AT AND POST-FAILURE
Degree of
Name Slide Classification
Slope Slide Dimensions Post-Failure Deformation Behaviour Break-up of
Slide Location Failure State Comments
(Hutchinson 1988) Width, Length, Depth, D Height Slope Volume Displacement Velocity Velocity the Slide
Geometry W/L W/D Comments
W (m) L (m) (m) (m) (deg) (cu.m.) (m) (mm/day) Category Mass
Limited movement, cracking at crest and bulging on mid
Park Reservoir Shallow slide in upper section of downstream slope likely slow
Rotational (?) Type 1 first time 12 21 5 11.5 33.7 750 0.57 2.40 Limited. - downstream slope. Velocity of movement likely to be slow Virtually intact Limited information.
Dam (above rockfill toe). Back-scarp at centre of crest. velocity
category.

Shallow slide in mid downstream slope. Crest of back-


Arroyito Dam Rotational scarp 7.5 m below crest on downstream slope. Very Type 2 first time 30 15 3 6 21.8 1000 2.00 10.00 No information. - unk No information on movement behaviour. unk -
small failure, possibly classify as a slough.

Slide in downstream slope with crest of back-scarp at 1000 to No information on movement behaviour. Suspect potentially
Type 1 or 2 4 to 5 2.1 m at crest of suspect suspect
Barton Dam Rotational (?) downstream edge of crest. Cracking in upstream side of first time 30 18 12.2 33.7 1500 1.67 6.67 unk moderate velocity given steepness of downstream slope and Limited information.
(?) (estimate) backscarp. moderate virtually intact.
crest. No information on slide depth. (estimate) potential for strain weakening of CH clay on shearing.

Estimate up to Slide observed on morning of 18 Dec. 1951. Crest had


Slide in downstream slope with crest of back-scarp at > 0.3 m at crest slow, possibly Toe of slide sand bagged and reservoir drawn down
1000 1000 (max.). moved 300 mm and toe 230 mm. The slide was still on the
Harrogate Dam Rotational (?) downstream edge of crest. No information on depth to Type 2 first time 24.5 16.2 4 to 6 (?) 8.8 27.8 1.51 4.90 and > 0.23 m at up to the low Virtually intact to stabilise the slope. Triggered by heavy and
(estimate) 300 next move at 12 to 13 mm/hour. Nearby slide developed as
surface of rupture. toe. end of moderate prolonged rainfall.
morning. bulging on lower slope followed by cracking at crest.

Woodrat Knot Slide in downstream slope. No further details beyond Movement occurred over a 3 month period. Started at a rate
Rotational (?) unk first time 90 unk unk unk 29.7 unk - - 9m 1920 moderate unk Limited information.
Dam this. of 80 mm/hr (1920 mm/day) and slowed gradually.

Suspect likely rapid velocity of movement given amount of Suspect


Slide coincident with highest stored water level in
Slide in downstream slope when reservoir at highest suspect movement and indication that movement had ceased. Further reasonable
Fruitgrowers Type 1 or 2 2000 to 2.7 m at crest, reservoir. Likely rapid velocity of initial failure due to
Rotational (?) historical level. Crest of back scarp at centre of crest. No first time 18 25 10 11 26.6 0.72 1.80 - moderate to movements were then observed (as well as water pouring amount of
Dam (?) 2500 7.6 m at toe. distance travelled. No indication that the failure
indication that the failure incorporated the foundation. rapid through cracks in back-scarp onto the slide mass) prior to break up at
incorporated the foundation.
intentional breaching of the embankment at the abutment. toe.

not known. Must Cracks in crest and along upstream slope to below stored Limited information given that total breach of
Deep-seated slide in downstream slope. Crest of back-
Santa Ana Compound (?) - possible be greater than water level observed in early November when reservoir at embankment followed the downslope failure.
scarp intersected the upper portion of the upstream 3000
Acaxochitlan translational base in Type 2 first time 20 35 11 12 26.6 0.57 1.82 about 1.5 to 2 m - suspect rapid 0.18 m below spillway. Failure occurred (in December) unk Occurred when reservoir at close to spillway level.
slope. Reservoir at close to spillway level when failure (estimate)
Dam foundation. for breach to several minutes after inspection by a watchman indicating This is close to the highest stored water level as the
occurred. Resulted in total breach of embankment.
occur. velocity of movement was likely to be in the rapid category. spillway had neven spilled before.

Toe and head of slide moved at a similar rate leading


Longitudinal crack observed on downstream face on 15/7/64
Compound (?) - Possible 1000 (avg max. up to and at the time of the large movement on the
Deep seated (for slope height) slide in downstream slope. when reservoir at full supply level. Monitoring shows a slow
translational component in 1.9 m at over 24 hours), 12/8/64. The monitoring shows a gradual increase in
Back-scarp at 1.6 m below crest level in downstream increase in the velocity of movement for 28 days before a very Minor break
Siburua Dam foundation and rotational Type 5 first time 24 14 3.5 6.5 26.6 600 1.71 6.86 backscarp, 1.7 m likely max rate moderate the rate of movement leading up the large movement
slope and toe of slide beyond toe of embankment. . large increase in velocity and large movement of the slide up.
backscarp according to at toe. of 2000 to on the 12/8/64. The reservoir was at full supply when
Reservoir at full supply level when slide initiated. occurred (on 12/8/64). Movement continued (at a much
vectors. 5000. the slide initiated and was in the process of being
slower rate) for the next 50 + days.
drawn down when the failure occurred.

13 failures occurred in 7 years. First indication is cracking on


Deep seated (for slope height) slide in downstream slope upstream side of crest (no vertical displacement). As
0.3 to 1.8 m at Failures occur where the field moisture content of the
Seven Sisters that passed through the foundation. Slide in section 3000 to movement progresses settlement occurs and cracks up to 0.3
Rotational (?) Type 5 first time 30 to 120 20 to 25 5 to 10 5 to 7.3 21.8 > 1.2 >3 crest, heave at - slow Virtually intact foundation is greater than 45% in the upper 2.4 m
Dike shows back-scarp of slide close to upstream edge of 15000 m wide develop over widths of 30 to 120 m. Over several
toe. (thus areas of lower undrained shear strength).
crest. weeks scarps of 0.3 to 1.8 m high develop with heave in toe
region.

Movement occurred 2 weeks after the reservoir had


Total movements reported as 3.35 m settlement (presumably
slow to reached the new maximum storage level. Potential
unk (but 60000 to of the crest) and 4.3 m displacement (presumably in the toe
Deep-seated slide in downstream slope that passed 3.35 m moderate (high risk of complete breach. Sandbagging on crest only
Compound (?), possibly likely to 80000 region). 4 days after slide started rate was 300 mm/day.
Great Western through the foundation. Slide back-scarp was below the Type 2 first time 180 unk ~ 18 m 18.6 >3 10.00 settlement, 4.3 m > 300 end of slow to Virtually intact accelerated movement. Drawdown narrowly averted
rotational. be about (rough Initial rate was likely to be higher than this, but not significantly
reservoir level on the upstream slope. displacement low end of breach (only 0.9 m of freeboard remained).
26) estimate) greater. Lesser movement on margins (3.05 m compared with
moderate) Placement of fill at downstream toe assisted in
4.3 m). Several scarps developed.
slowing slide movement.

Likely rapid failure. Occurred 10 months after


80000 to 10 m at crest, 9 Failure occurred one morning in April 1978. "Slipped with a completion of construction. Reservoir had been at full
Compound (?) - likely basal Deep-seated slide in downstream slope that passed
Type 5 or 2 100000 Likely
Aran Dam translational component through the foundation. Slide back-scarp was at the first time 150 75 17 21 23 2.00 8.82 to 11 m (maybe unk likely rapid rumbling noise". Suspect movement occurred at a rapid supply level for some 8 months. Rumbling noise
(?) (rough significant
through foundation. downstream edge of the crest. greater) at toe. velocity. suggests internal shearing, possibly on lateral
estimate)
margins.

Slide in downstream slope in section of embankment


where the height was about 4.6 m. Top of back-scarp Only reference to movement is the total movement at the top
Lake Yosemite Rotational (?), maybe 0.6 m at crest of slow to
was on the upper downstream slope, 1.5 m below crest Type 2 first time 45 6 2 3.1 26.6 300 7.50 22.50 unk of the backscarp. No indication of time period of movement. Intact (?) Limited information.
Dam compound (?) backscarp. moderate
level. Sherard (1954) refers to the slide surface passing Suspect slow to moderate given limited amount of movement.
through the upper 0.5 m of the foundation.

Compound - basal Slide in downstream slope that passed through the upper 1.5 m at crest, 11 Triggered by earthquake. Slide mass reported as
rapid to very Indications that slide occurred within minutes according to
Yuba Dam translational component in (less than 0.5 m) portion of the foundation.. Top of back- Type 2 first time 23 16 4.5 7.6 29.7 1000 1.44 5.11 m at toe (from unk Significant "completely fissured and loose" indicating a
rapid reports, therefore very likely at rapid to very rapid velocity.
foundation. scarp at the centre of the crest. section) significant degree of break up during failure.
Appendix D Page D21

Table D1.3: Case studies of slope instability of embankment dams – post-construction slides in the downstream slope (Sheet 4 of 4).
HYDROGEOLOGY MECHANICS OF FAILURE
Name Other Comments References
Rainfall Seepage Comments Cause/s Landsliding Trigger Failure Mechanism Comments on Mechanism

Possible drained failure through dilative silty and clayey sand earthfill. Small volume slide through the embankment only.
Park Reservoir Possible seepage on downstream face associated with
No information. No information Likely seepage or rainfall. Likely seepage or rainfall. Shallow failure (therefore low confining pressures). Limited movement Very limited information available on slide. Embankment constructed of silty and clayey sands. USCOLD (1975)
Dam layering in earthfill (kh >> kv), but no information.
on steep slope (1.5H to 1V). Very limited information.

Perched water tables above phreatic surface Reservoir level raised to operating level in 1982. Saturation of fill, perched High localised phreatic Drained failure in sandy dilative sandy gravel fill with low (< 10%) fines Small volume slide through the embankment only.
Giuliani and Pujol
Arroyito Dam No information. due to stratification. Seepage observed on Seepage observed on downstream face. Top of water table/s in embankment surface and seepage. content. Possibly a sloughing type failure due to seepage pressures in - Embankment constructed of well-compacted sandy
(1985)
downstream face. backscarp is below reservoir operating level. and seepage. Saturation of fill. cohesionless fill. Small volume and shallow depth. gravels. Very limited information.

Seepage appeared on downstream slope


Reservoir close to full supply level when slide occurred Small volume, likely shallow slide, through
directly after construction. Increased in area Initiated as a drained failure. Likely progressive weakening of fill due to Eventually reached a state of limiting stability, at which point failure
(had been for at least several months, possibly a lot Seepage. Gradual saturation Seepage and saturation of embankment. Embankment constructed of CH
Barton Dam No information. with time until slope saturated to mid-height. gradual saturation and reduction of matrix suction. Possible strain occurred. Steepness of downstream slope contributed to amount of Sherard (1953)
longer). Likely stratification in embankment (kh > k v) of downstream slope. downstream slope. lacustrine clays (without filters) with no formal
May 1922 - considerable moisture on weakening on shearing in CH clays to residual strength. movement.
and high phreatic surface. compaction. Steep downstream slope.
downstream slope.

Annual average = 359 mm. Initiated as a drained failure in a water softened and saturated Small volume failure in relatively steep downstream
1951 - 656 mm for year Failure occurred in the winter of a very wet year and Rainfall. Filling of deep downstream slope (softened and saturated by water infiltration and Surface of rupture on major slip plane was about 12 mm thick and had the embankment slope of medium plasticity clays.
and 210 mm in Nov, was preceeded by a very wet November (wettest shrinkage cracks and exaserbated by deep shrinkage cracks). Likely high phreatic surface or consistency of "porridge" (soft to very soft ?) clay. Limiting stability Triggered by period of heavy and prolonged
Harrogate Dam No information Rainfall. Davies (1953)
highest yearly and monthly month on record). Water infilled shrinkage cracks and softening of fill in downstream perched water table/s (due to rainfall). Slope at limiting stability under condition in drained loading for relatively steep slope in medium plasticity rainfall. Numerous shrinkage cracks in
(Nov) rainfall since wetting up of downstream shoulder from rainfall. shoulder. these conditions. Possibly some strain weakening on shearing to clays. embankment contributed to saturation of slope and
embankment constructed. residual strength. instability.

Likely progressive weakening of fill due to gradual saturation and


After 3 years operation dampness observed
Gradual saturation of downstream slope with time reduction of matrix suction. Initiated as a drained failure once a limiting Eventually reached a state of limiting stability, at which point failure Suspect failure through embankment only. Limited
Woodrat Knot on downstream slope, about 11 m above the Seepage. Gradual saturation Seepage and saturation of
No information. leading up to failure. No information on reservoir stability condition was reached in the relatively steep downstream occurred. Steepness of downstream slope contributed to amount of information on material properties but reasonable ICOLD (1974)
Dam toe. Over the next 2 years the extent of the of downstream slope. downstream slope.
operation. slope. Likely strain weakening on shearing resulted in large movement. Large movements indicate strain weakening on shearing. information on slide movement.
dampness increased.
movements at moderate velocity.

Initiated as a drained failure due to limiting stability of the downstream


Reservoir reached highest level in embankment history
Sloughing prior to 1930's indicative of slope. Wetting up and softening of the embankment filling, a high
(2.1 m below crest level) 6 days prior to failure. High phreatic surface. Rise of the reservoir level to
seepage through the embankment. Stream phreatic surface and increased seepage pressures (all associated with No definite indication that the failure occurred through the foundation. It is Failure in downstream slope of embankment.
Fruitgrowers Observations indicate a high phreatic surface through Wetting up and softening of the within 2.1 m of crest level, Sherard (1953)
No information. of water poured from slide area about 4.3 m highest water level) all contributed to the failure. Large amount and possible that the slide occurred in the embankment only and travelled a Occurred 6 days after reservoir reached highest
Dam the embankment. Possible shrinkage cracking the downstream slope. highest level in embankment ENR (1937b, 1939b)
below crest level (coincident with crest level likely rapid velocity of movement in indicative of strain weakening on sufficient distance in the toe region to take out the outlet works. stored level. Likely rapid failure.
associated with staged construction may have Seepage pressures. history.
prior to 1905 raise). shearing. Possibly due to dilation of slide mass on shearing and
contributed to formation of seepage paths.
availability of water to further soften the slide mass.

Slide coincident with high stored water level in Adverse change in seepage Slide brittleness potentially due to a number of factors including: restraint
Initiated as a drained failure. An increase in the phreatic surface, Downstream slope failure that extended back to the
Santa Ana Moisture observed on downstream slope in embankment. Cracking in early November resulted in High reservoir level. Cracking path conditions resulting in a on the lateral margins of a relatively narrow slide (W/L = 0.6); strain
wetting up and softening of the downstream slope and increased upstream slope and resulted in the breach of the Marsal and Tamez
Acaxochitlan No information. early to mid November 1952 (shortly after a change in the seepage path conditions through the and adverse change in higher phreatic surface, and weakening on shearing of the high plasticity embankment materials and
seepage pressures all contributed to the failure. Potential rapid velocity embankment. Likely rapid failure within the (1955)
Dam cracking observed). embankment and moisture observation on the seepage path conditions. wetting up and softening of foundation; and entry of water into the dilating slide mass as failure is
of movement suggests brittleness in the failure mechanism. embankment and foundation.
downstream slope. the downstream shoulder. occurring resulting in further softening.

Suspect the failure may be driven by the foundation, along a possible relict
Likely low phreatic surface in embankment fill.
bedding joint (bedding favourable oriented in region of slide). High
Piezometers in foundation show phreatic surface Initiated as a drained failure whena reaching a limiting stability
moisture contents in the slide area indicate preferential wetting up in this
Significant increase in flow rate in drain at toe maintained within the foundation. High swelling Gradual wetting up and Gradual wetting up and condition reached in the downstream slope. Possible progressive Small failure through homogeneous embankment
area, possibly due to localised seepage path in the embankment (?). Filter
of slope (in the area of the slide) when movement in the vicinity of the slide may indicate softening of downstream softening of downstream failure in potentially dilative embankment materials (at low confining and foundation of medium plasticity shaly clays.
material decribed as poor quality and containing 40% fines, but still flowed Wolfskill and Lambe
Siburua Dam Relatively low rainfall area. reservoir at full supply level. Indicative of seepage occurred through the fill and resulted in the embankment and foundation. embankment and foundation. pressures and non saturated) and heavily over-consolidated Slow increase in rate of movement up to failure
at high reservoir level. The foundation slope (7 to 11 degrees downstream) (1967)
likely seepage through the foundation given wetting up and swelling of the embankment. Lower, Initial cracking occurred when Initial cracking occurred when foundation. Rapid velocity of movement suggests strain weakening on (over 28 days) and then continued movement once
may also be a factor in the failure. Possible progressive failure, within the
piezometric levels. weathered foundation responds with reservoir level. reservoir at full supply level. reservoir at full supply level. shearing of medium plasticity shaly clays and/or restraint on the lateral failure occurred for some 50 days.
foundation, of a slope under relatively high shear stress conditions. Back
Upper foundation piezometer indicates a very limited margins.
analysis gives drained failure at less than peak strength (assuming
response to reservoir level.
embankment near saturated).

Undrained failures through low undrained strength foundation and wet


No information. Suspect development of phreatic
Possible rise in phreatic placed, low undrained strength CH earthfill. For failures more than 6 to Numerous failures in downstream slope after
Seven Sisters surface in foundation and/or embankment enough to Rise in phreatic surface in Would consider CH foundation and wet placed CH earthfill are potentially
No information. No information surface in embankment or 12 months after end of construction suspect the development of the construction. Typical of slow failures of Peterson et al (1957)
Dike trigger failure. Potential seepage path along cracked embankment or foundation. strain weakening in undrained loading.
foundation. phreatic surface had an influence on stability (i.e., enough to trigger embankments on soft ground.
and fissured zone of foundation.
slide in slope of limiting stability in undrained conditions).

Initiated as a drained failure due to limiting stability of the downstream Very limited information on material properties. If the foundation is driving
Downstream slope failure that extended back to
slope under an increasing phreatic surface, pore pressures and/or the slide then suspect that spreading of the slide width occurred due to
High water level in reservoir below the reservoir level on the upstream slope.
Raising of embankment, high seepage pressures associated with the new high reservoir level. restraint on the lateral margins where shearing through the embankment
Movement started some 2 weeks after reservoir level and associated change in the Very close to a breach condition of the Sherard et al (1963)
reservoir water level and Occurred 9 months after raising embankment some 4 m and 2 weeks was required. In addition, the lateral persistence of the clay foundation
Great Western No information. No indication. raised to new maximum storage level. Suspact phreatic surface or pore embankment, which may have only been averted ENR (1958)
increasing phreatic surface after reservoir reached new maximum storage level. Suspect initiated may have been a factor in the slow velocity of movement. Other factors
increase in phreatic surface enough to initiate failure. pressures in zones of the because of the relatively low velocity of movement. ICOLD (1974)
through the embankment. in the alluvial clay foundation (referred to a "unstable" and "weak"). could also have contributed to the slide such as permeability differential
embankment or foundation. Failure occurred within the embankment and
Possible strain weakening on shearing and possible restraint on lateral between the new and old fill, and the presence of shrinkage cracks in the
foundation.
margins. old embankment that were inundated when the embankment was raised.

Suspect initiated as drained instability (in foundation) and possibly


Reservoir at full supply level since August 1977 (some Very limited information from which to draw any firm conclusions as to the
A number of possibilities. driven by gradual strength reduction of foundation on wetting (i.e., Limited information. Relatively large failure in
8 months prior to failure). Phreatic surface (shown by mechanism of failure. If restraint on the lateral margins was significant it
Rising pore pressures in reduction of suction in the foundation had a significant effect on downstream slope that occurred 9 months after end
Dighe et al, 1985) shows relatively high level in would be expected that pre-failure signs would have been clearly evident
Aran Dam No information. No information foundation and downstream Not sure. stability). Rapid movement suggests brittleness of failure mechanism, of construction and some 8 months after reservoir Dighe et al (1985)
downstream shoulder fill. This would suggest the for a slide of 150 m width. Possible graben type failure that moved rapidly
shoulder, softening of possibly on the lateral margins but maybe internally as well. Suspect at full supply level. Within embankment and
downstream shoulder fill is of relatively low once internal shearing through the downstream shoulder fill (rumbling
foundation on wetting. the foundation may also strain weakening on shearing, but that this is foundation.
permeability. noise) occurred and released the slide.
not the dominant cause of brittleness.

Seepage observations indicate the foundation is likely


Limited seepage had always occurred under Not known. Possibly localised
to be of greater permeability than the embankment. Limited information. Possible change in seepage conditions resulted in Localised small volume failure in low height section
the embankment and had formed marshy change in seepage conditions Drained instability in downstream slope. Not a sloughing type failure.
Lake Yosemite Likely relatively high phreatic surface in the localised instability of downstream slope or heavy rainfall triggered the of the embankment that occurred some 60 years
No information. areas at the downstream toe. On occasion Not known. through the foundation or Reports during repair indicate the surface of rupture passed through Sherard (1953)
Dam embankment. For such a localised failure to occur so slide. Dilation on shearing and availability of water may have resulted in after initial construction. Incorporated the
the bottom metre of the embankment slope embankment. Possibly the foundation.
long after construction may suggest heavy rainfall further softening. foundation.
had been reported as moist. triggered by rainfall.
preceeded the slide.

Prior to raise in 1949, considerable moisture High phreatic surface developed within the Limiting stability condition of the downstream slope prior to the earthquake Relatively small volume failure triggered by
was observed on the downstream slope. embankment due to the presence of the low Earthquake. High phreatic Earthquake was sufficient to trigger the slide. Likely that the slope was and landslide. Likely that earthquake sufficient to reduce the factor of earthquake. Rapid to very rapid failure within the
Yuba Dam No information. After the raise no moisture was noted permeability layer capping the downstream slope and surface due to permeability Earthquake. of limiting stability prior to earthquake given the high phreatic surface safety to less than one. The relatively large movement suggest strain embankment and foundation. Went close to Sherard (1953)
indicating new fill material of lower crest. Likely increased pore pressures within the differential. and relatively steep slope in medium plasticity sandy clays. weakening on shearing. Possible strain weakening in the outer breaching the embankment (backscarp at centre of
permeability than original material. foundation. downstream zone and foundation. the crest).
Appendix D Page D22

2.0 FAILURES IN CUT SLOPES IN HEAVILY OVER-


CONSOLIDATED HIGH PLASTICITY CLAYS

2.1 CUT S LOPES IN LONDON CLAY

The analysis of the post-failure deformation behaviour of cut slopes in the heavily over-
consolidated high plasticity London clay comprises twenty-five case studies, the details
for which are summarised in Table D2.3. The table provides details for each case study
on; timing of the failure, slope type, slope geometry, material properties, slide
geometry, post-failure deformation behaviour, hydro-geological conditions and
references.
The London clay formation (Skempton et al 1969) is a marine clay of Eocene (early
Tertiary) age. Following uplifting it has been subjected to erosion in the late Tertiary
and Pleistocene periods, with estimated erosion depths of 150 to 300 m (Skempton et al
1969). The formation is heavily over-consolidated as a result of the significant depth of
erosion and importantly, in terms of the assessment of cut slope instability, it is of
uniform lithology in the London area (Skempton et al 1969).
The typical profile of the London clay formation (Skempton et al 1969; Skempton
1977) comprises a weathered and oxidised upper zone of 5 to 15 m depth, termed
“brown” London clay by virtue of its colour, underlain by a thin transition zone inturn
underlain by the unweathered “blue” London clay. The typical properties of the London
clay (both brown and blue) are summarised in Table D2.1. It is a high plasticity
fissured clay of low permeability and typically of stiff to very stiff undrained strength
consistency.
The discontinuities and defects within the upper part of the London clay formation
(Skempton et al 1969) are summarised as follows:
• Bedding – near horizontal and typically observed as a gently undulating surface.
Not easily discernable as there is generally no lithological change.
• Jointing – typically a near vertical set of orthogonal joints with a thin layer of gouge
clay on a plane surface. Sheeting joints are also observed with dips of between 5
and 15 degrees.
• Fissuring – In the upper 10 to 15 m the fissures are typically less than 150 mm in
size with no preferred orientation, but with a tendency to concentrate in the sub-
horizontal direction parallel to the bedding. The number of fissures increases and
Appendix D Page D23

the size decreases as the ground surface is approached. Typically, no appreciable


relative movements have occurred along the fissures and the fissure surface has a
matt texture. However, about 5% of the fissure surfaces are slickensided and have
polished and striated surfaces.
• Faulting – limited.

The strength properties of London clay indicate that it is strain weakening in both
drained (Figure D2.1) and undrained loading (Figure D2.2). The typical drained
strength properties (Table D2.1) show the strength on defects is less than that of the
intact soil.

Table D2.1: Summary of material properties of London clay*1


Material Property “Brown” London Clay “Blue” London Clay
Unit Weight (kN/m3) - 18.9

Plasticity
Liquid Limit (%) 82 (62 to 95)
Plastic Limit (%) 30
Grading
Clay fraction (%)*2 55 to 60
Drained Strength
Peak c′ = 31 kPa, φ′ = 20° c′ = 18 (15 to 23) kPa , φ′ = 20° (19 to 21)
Fully softened c′ = 0 to 1.5 kPa, φ′ = 20° c′ = 0 to 1.5 kPa , φ′ = 20°
Residual c′ = 0 to 1.5 kPa, φ′ = 16° c′ = 0 to 1.5 kPa , φ′ = 13° to 16°
Joints and Fissures c′ = 7 kPa , φ′ = 18.5° -
3 Stiff increasing to very stiff Very stiff
Undrained Strength*
with depth
Permeability - 2 to 4 x 10 -9 m/sec
*1 Referenced from Skempton (1964, 1977), Skempton and La Rochelle (1965), Skempton et al (1969).
*2 The clay fraction is as that fraction less than 2 micron.
*3 Undrained strength terminology to Australian Standard AS 1726-1993 “Geotechnical Site
Investigations”. Stiff, Su = >50 and ≤ 100 kPa; Very Stiff, Su = >100 and ≤ 200 kPa, where Su =
undrained shear strength.
Appendix D Page D24

Figure D2.1: Drained strength properties of ‘blue’ London clay at Wraysbury


(Skempton et al 1969)

Figure D2.2: Undrained strength properties of London clay, Bradwell (Skempton and
La Rochelle 1965)

2.2 CUT S LOPES IN WEATHERED UPPER LIAS CLAY

The analysis of the post-failure deformation behaviour of cut slopes in the heavily over-
consolidated, weathered Upper Lias clay comprised 12 case studies, the details for
which are summarised in Table D2.4. The table provides details for each case study on;
Appendix D Page D25

timing of the failure, slope type, slope geometry, material properties, slide geometry,
post-failure deformation behaviour, hydro-geological conditions and references.
The Upper Lias clay formation (Chandler 1972) is a marine clay-shale of Jurassic
age. Following uplifting the formation has been subjected to erosion, estimated at
depths of up to 600 m (Chandler 1972), resulting in its present heavily over-
consolidated condition. Chandler (1984a) indicates that the formation presently
outcrops across the UK from the north eastern coast (south of the English – Scottish
border) to the central region of the southern coast. The case studies of cut slope failures
are located in the midlands region where extensive outcrops of the Upper Lias occur
(Chandler 1972).
The Upper Lias clay is of relatively uniform lithology, although two different upper
surface profiles have been identified by Chandler (1972) that are dependent on the
development history of the slope during the Quaternary age. Chandler indicates that
during the Pleistocene age the upper profile was significantly disturbed from permafrost
conditions. He describes the disturbed, or brecciated, zone as “small stratified lumps or
lithorelicts in a matrix of softer clay” and indicates the depth of brecciation at the time
of disturbance was up to 10 to 11 m. Chandler (1972) found this brecciated profile to
still be evident at some outcrop locations of the Upper Lias clay, and termed this the
‘brecciated’ profile. At other outcrops he identified that the zone of brecciation had
been eroded and termed this the ‘fissured’ profile.
The material properties of the brecciated and the fissured weathered Upper Lias clay
are summarised in Table D2.2. The weathered formation is of high plasticity and high
clay fines content, and typically of stiff to very stiff undrained strength consistency.
Chandler (1972) comments that the brecciated profile is typified by a lack of fissuring
and a permeability one to two orders of magnitude greater than for the fissured profile.
For the fissured profile he indicates that the fissuring is typically limited to the upper
weathered zone and does not extend into the underlying un-weathered clay-shale
profile. Other defects within the soil mass include near horizontal bedding.
Slickensiding is observed on fissures in the more weathered materials.
The strength properties of the weathered Upper Lias clay are similar to London clay
in that the soils are strain weakening in both drained and undrained loading. It is also
suspected that the drained strength properties on fissures and other defects are likely to
be less than that of the intact soil, and have a significant effect on the strength of the
overall soil mass.
Appendix D Page D26

Table D2.2: Summary of material properties of weathered Upper Lias clay*1


Material Property ‘brecciated’ clay ‘fissured’ clay
Unit Weight (kN/m3) 19.0 19.2
Plasticity
- Liquid Limit (%) 60
- Plastic Limit (%) 28
- Plasticity Index (%) 32
Grading
Clay fraction 45
Drained Strength
- Peak c′ = 17 kPa, φ′ = 23° -
- Fully softened c′ = 1 kPa, φ′ = 23° c′ = 1 to 2 kPa , φ′ = 25°
- Residual * 2
c′ = 0 kPa , φ′ = 10 to 17.5° c′ = 0 kPa , φ′ = 10 to 17.5°
Undrained Strength*3 Stiff to very stiff increasing to Stiff to very stiff, increasing with
hard with depth depth
-9 -10 -10
Permeability 10 to 10 m/sec Less than 10 to as low as 10-12
m/sec
1
* – Referenced from Chandler (1972, 1974, 1976)
*2 – Dependency of residual strength on effective normal stress (refer Figure D2.3).
*3 Undrained strength terminology to Australian Standard AS 1726-1993 “Geotechnical Site
Investigations”. Stiff, Su = >50 and ≤ 100 kPa; Very Stiff, Su = >100 and ≤ 200 kPa; Hard, Su =
>200 kPa, where Su = undrained shear strength.

Figure D2.3: Effect of the normal effective stress on the residual strength of weathered
Upper Lias clay (Chandler 1976).

2.3 TERMINOLOGY USED IN TABLES


An explanation of some of the terminology used in Table D2.3 and Table D2.4 is as
follows:
• Some general terms or symbols used and their meaning are:
Appendix D Page D27

− (?) = unsure of detail, but thought to be likely


− unk = unknown
− “-” = not required, not appropriate or not given
− est = estimate
− H to V refers to horizontal to vertical slope angle; i.e. 2H to 1V.
− Su = undrained shear strength
− Soil classification symbols, where used, are in accordance with the Australian
soil classification system (Australian Standard AS 1726-1993 “Geotechnical
Site Investigations”).
• Under the section on slope geometry;
− For slope type, “Cut” = cut slope, “Cut-R” = retained cut slope. Only those
failures in retained slopes where the surface of rupture passed below the
retaining wall are classified “Cut-R”.
− The slope angle given for the partially retained cuts is the slope angle of the
unsupported section of the cut, i.e. the slope angle above the retaining wall.
• For the material properties:
− Undrained shear strength terms used are in accordance with the Australian
soil classification system. Very Soft ≤ 12 kPa; Soft > 12 and ≤ 25 kPa; Firm
> 25 and ≤ 50 kPa; Stiff > 50 and ≤ 100 kPa; Very Stiff > 100 and ≤ 200 kPa;
Hard > 200 kPa.
• “Slide classification” refers to the classification of the shape of the surface of
rupture (e.g., translational, rotational or compound) using the landslide classification
system of Hutchinson (1988).
• “Slope Failure Geometry” refers to the location of the slide within the slope. Details
are given in Section 1.3.2 of Chapter 1 and in Appendix A.
• For the slide dimensions, the slope angle allows for benching or partial retainment
of the slope and it will, in some cases, be slightly different from the cut slope angle
reported under the section under slope geometry. Only a small number of the case
studies are affected by this, mainly partly retained cuts, otherwise the slope angle is
equivalent to the cut slope angle.
• The limit equilibrium analysis refers to the drained strength parameters from back-
analysis by others at failure (i.e. a factor of safety of 1). For the London clay case
studies φ′ has been held constant at 20° for first time failures but is reported at less
Appendix D Page D28

than 20° for reactivated slides. The same principle has been applied to the cut
slopes in Upper Lias clay with φ′ held constant at 25° for the ‘fissured’ profile and
23° for the ‘brecciated’ profile.
• For the post-failure deformation behaviour:
− For the deformation column, “head” and “toe” refer to the head and toe
regions of the slide.
− Velocity category refers to the IUGS (1995) velocity categories.
− The degree of break-up is a somewhat subjective assessment of the post-
failure break up of the slide mass.
Appendix D Page D29

Table D2.3: Case studies of failures in cut slopes in London clay


Slope Geometry Material Properties Slide Details
Age of
Date of Slope at Pre-Failure Slope Slide Dimensions
Name Slope % of Slope
Failure Failure Slope Height, Plastic Limit Undrained Deformation State of
Angle Height Comments Liq Limit (%) Slide Classification Failure
(years) Type H s (m) (%) Strength Sliding Width Height, Depth, D Slope Volume
(Degrees) Retained Geometry
(m) H f (m) (m) ( o) (cu.m.)

6 to 6.5 mm/year in 1929.


Large retaining wall at toe of cut. Cut Increase in rate after 1936 to Compound first time Type 5 unk 8.2 5.8 29.7 unk
Kensal Green 1941 29 Cut-R 6.1 15 61% - - -
slope above retaining wall at 15 degrees.
1937 up to failure in 1941.

Cut retained at toe of slope. Cut slope Compound - translational


tension cracks observed
Upper Holloway 1951 81 Cut-R 5.3 22.2 20% above retaining wall at 22.2 degrees. Flat 86 28 - base, rotational toe and first time Type 5 unk 6.25 5.2 25 unk
prior to failure
slope at top of cut. back-scarp

Capped by layer of sand and gravel. Flat


Fareham 1961 59 Cut 9.5 19.5 - - - - na rotational reactivated Type 2 > 60 12 6.5 19.5 8 to 10,000
slope above cut.

West Acton - A 1966 44 Cut 4.7 18.4 - Flat slope above cut. - - - na rotational (?) first time Type 1 unk 4.1 2.4 18.4 unk

West Acton - B 1966 44 Cut 4.8 18.4 - Flat slope above cut. - - - na rotational (?) first time Type 2 unk 4.6 2.2 18.4 unk

Flat slope above cut. Drain at toe of Compound - translational


Grove Park 1962 98 Cut 9.4 22.5 - - - - na first time Type 1 unk 4.8 3 22.5 unk
slope base, rotational back-scarp

Capped by layer of sand and gravel. Flat


Dedham 1952 112 Cut 10.2 22 - - - - na rotational (?) First time Type 2 15 8.5 3.6 22 600
slope above cut.

Drain at crest of slope. Flat slope above na rotational first time Type 2 unk 6 2.3 23.5 unk
Kingbury - S1 1947 16 Cut 6.2 23.5 - - - -
cut.
Movement noted during
Counterfort drains at 45 m spacings. Compound - translational first time (possible
St. Helier 1952 22 Cut 7.3 25.3 - - - - periods of wet weather in Type 2 < 45 7.6 4.7 25.3 600 to 1500
Capped by layer of sand. toe, rotational back-scarp reactivation)
years prior to slide in 1952.

Crews Hill 1956 47 Cut 5.3 16.2 - Capped by layer of gravel and sand. - - - na Translational (shallow) first time Type 2 unk 5.3 1.1 16.2 unk

Compound - translational
stiff (very soft on na basal component, reactivated Type 2 unk 9 4 20.5 unk
Althorne 1957, October 60 Cut 8.8 20.5 - Flat slope at crest. - -
surface of rupture)
rotational back-scarp
softened zone
Translational (shallow at
Cuffley 1951 (winter) 38 Cut 12 20.7 - Flat slope at crest. - - defining surface of na first time Type 2 unk 12 4 20.7 unk
toe)
rupture

1947 - slope trimmed back from 3.5H:1V stiff, disturbed is soft Compound - significant first time (possible
Hadley Wood 1951 70 Cut 13.6 13.5 - - - na Type 1 unk 10.4 2.2 13.5 unk
to 4H:1V. (and soggy) translational component reactivation)

Compound - translational
Small retaining wall at toe (not involved in
Whitstable 1959 99 Cut 11.4 18.4 13% - - - na on brown/blue clay first time Type 2 unk 9.9 4.4 18.4 unk
slip). Flat slope above cut.
interface.

Small retaining wall at toe (not involved in na Rotational / Compound (?) first time Type 1 unk 13 6 17.8 unk
Grange Hill 1950 50 (?) Cut 17.4 17.8 14% - - -
slip).

Isle of Sheppey - A na 6 Cut 12 38 - Steep cut. - - soft to firm (?) na rotational (?) first time Type 2 unk 12 4.5 38 unk

Isle of Sheppey - B na 8 Cut 10 41 - Steep cut. - - soft to firm (?) na rotational (?) first time Type 2 unk 10.1 3.5 41 unk

Steep cut in London clay. Flatter


benched slope in upper region of slope Bulging and cracking of clay
Bradwell 24-Apr-1957 0.014 Cut 14.8 63 - 95 30 firm to very stiff Rotational, retrogressive first time Type 2 27 14.8 9 56 2500
(average 29 degrees) in filling and Marsh face in toe region.
Clay.
Large retaining wall at toe of cut. Cut
1948, stiff to very stiff (slip
Wood Green 55 Cut-R 11.3 18.4 43% slope above retaining wall at 18.4 78 30 na rotational (?) first time Type 5 73 14.4 8 28 7 to 10000
November surface - firm).
degrees.
Compound - translational
Sudbury Hill 1949 46 Cut 7 18.4 - Flat slope above crest of cut. 82 28 - na first time Type 2 unk 7 4 18.4 unk
toe, rotational back-scarp

Initial slow creep type


Compound - translational
New Cross 2-Nov-1841 3 Cut 23 33.7 - Steep cut. - - - movements observed on first time Type 1 110 17 13 33.7 40000
toe, rotational back-scarp
day of slide.

Small retaining wall at toe (not involved in stiff to very stiff (slip Tension cracks at crest of Compound - translational
Northolt 1955, January 19 Cut 10.1 21.8 7% 78 28 first time Type 1 83 6.8 4.4 21.8 3000
slip). surface - soft). slope toe, rotational back-scarp

1918, Large retaining wall at toe of cut. Slope Movement observed several
Wembley Hill 13 Cut-R 17 18.5 41% - - - Compound first time Type 5 260 19.5 10.5 26 70000
February above retaining wall cut at 18.5 degrees. days prior to failure.

Large retaining wall at toe of cut. Slope Observations of cracking


1954, stiff to very stiff (slip Compound (?) - rotational
Uxbridge - 1954 54 Cut-R 10.2 18 77% cut at 18 degrees above the retaining 86 24 and movement at crest and first time Type 5 85 13.3 12 32 8 to 10000
November zone soft to firm) behind retaining wall
wall. toe of slope from Sept 1954.

5 m wide excavation for new retaining


stiff to very stiff (slip
Uxbridge - 1937 1937 37 Cut 9 26 - wall behind existing retaining wall. 86 24 na Rotational first time Type 2 32 9 4.6 43 2500
zone soft to firm)
Vertical cut, slope above = 26 degrees.
Appendix D Page D30

Table D2.3: Case studies of failures in cut slopes in London clay (Sheet 2 of 2)
At And Post-Failure Hydrology
Limit Equilibrium Analysis
Post-Failure Deformation Behaviour
Name (from references) Rainfall / Cause/s of Sliding Trigger Comments References
Hydro-geological Features
Deformation Peak Velocity Velocity Degree of Snowmelt
c' (kPa) phi ' (deg) Comment
(m) (mm/day) Category Break-up

Tension crack observed 6 m behind wall and thw wall had slid
High seasonal pore Initial excavation in 1836. Widened and retained in 1912. Earlier
Kensal Green 4 20 0.3 20 to 100 slow Intact forward by 300 mm. Continued movement of 100 mm over Phreatic surface assumed unk Excavation Skempton (1964)
pressures (?) failure in 1927. This section was first time failure.
following 8 months.

1.3 (crest), 0.2 to Initial movements observed in 1951. New wall built in 1953 also Piezos installed in 1956 and records High seasonal pore Continued movement after initial slip in 1951. Failure at fully softened James (1970)
Upper Holloway 0.5 20 50 to 200 (est.) suspect slow Intact unk Excavation
0.7 (toe). failed ( >300 mm movement between July 1953 and Apr 1955). used to estimate phreatic surface. pressures (?) strength. Suspect slow to moderate initial rate. Lo and Lee (1973)

Initial slip in winter of 1961/62 with approx. 0.5 to 1 m movemetn Several large failures during construction therefore slopes cut back
1.5 (crest), 2 to 4 Phreatic surface estimated from High seasonal pore
Fareham 0 18 to 19.5 200 to 400 (est.) suspect slow partly broken up at toe (possible reactivation). Further movements in winter of unk Excavation and drains installed. Suspect reactivated slide, therefore slow seasonal James (1970)
(toe) observations pressures (?)
1967, up to 2.5 to 3 m at toe (definite reactivation). creep movement.

0.36 (crest), 0.9 to No details on timing of movement, but suspect slow given limited High seasonal pore
West Acton - A 0.5 20 200 to 400 (est.) suspect slow virtually intact Phreatic surface assumed unk Excavation First time failure. James (1970)
1.5 (toe) movement of slide. pressures
0.41 (crest), 1.5 to No details on timing of movement, but suspect slow given limited High seasonal pore
West Acton - B 1 20 200 to 400 (est.) suspect slow virtually intact Phreatic surface assumed unk Excavation First time failure. James (1970)
2.5 (toe) movement of slide. pressures
Slide occurred in 1962. No indication of time over which
Piezos indicate slip surface above Failure in upper section of cut (Type 1). Translational component on
3.3 (crest), 10.8 50000 to 400000 movement occurred. Suspect moderate at crest and rapid at toe. High seasonal pore
Grove Park 0.5 20 rapid (toe) broken up phreatic surface. Possibly high unk Excavation bedding plane. Travel on 22.5 deg. cut slope angle, at likely rapid James (1970)
(toe) (est.) Type 1 failure with 10.8 m movement beyond toe on 2H to 1V pressures (?)
phreatic surface at time of slip ? velocity.
slope.
Initial slip in 1910/11 (no details). Further slip in 1952, moved up Large slip in winter of 1910/11, stabilised by drainage trenches. 1952
2.1 (crest), 3.5 to 5 Phreatic surface from water levels in High seasonal pore
Dedham 2.5 20 500 to 1500 (est) moderate partially broken up to 3.5 to 4 m at toe. When toe cut back movement continued for unk Excavation slip located near to the 1910 slip. Continued movement for weeks James (1970)
(toe) standpipes after failure. pressures (?)
up to 6 weeks. after initial slide suggests initially slow movement.
0.5 (crest), 0.5 to 1 No details on timing of movement. Suspect slow given small High seasonal pore After slip slope was trimmed back to 2.5 - 3H to 1V and conterfort
Kingbury - S1 3.5 20 50 to 200 (est.) slow virtually intact Phreatic surface assumed unk Excavation James (1970)
(toe) deformation. pressures (?) drainage was installed.
Slope trimmed and drains installed in 1948 (possible previous sliding).
Initial failure prior to 1948 when slope trimmed back and drains Phreatic surface from water levels in
High seasonal pore 1952 slip between to counterfort drains. Movement noted prior to 1952
St. Helier 4.5 20 2.3 (crest), 2.5 (toe). 200 to 600 (est.) slow to moderate Minor break up installed. From 1948 movement observed between counterfort standpipes. Movement occurred Wet weather Excavation James (1970)
pressures (?) during wet weather. Slipped material described as soggy with
drains. Suspect 1952 movement is a reactivation. during wet weather.
moisture content close to liquid limit.

Shallow translational slip possibly on discontinuity. Movement prior to


0.5 to 1.2 (crest), 1 Suspect possible prior movement of slip, and therefore 1956 High seasonal pore
Crews Hill 1 20 100 to 400 (est.) suspect slow partially broken up Phreatic surface assumed unk Excavation 1956 had been noted. Slipped clay described as soft and soggy. James (1970)
to 1.5 (toe) movement is a reactivation. pressures (?)
Suspect possible reactivation and r u is greater than 0.3.

1.5 to 2 (crest), 3 m Suspected reactivation of previous slip. Still active at time of Phreatic surface assumed, High seasonal pore
Althorne 0 16.5 200 to 400 (est.) suspect slow partially broken up unk Excavation Date of initial slide not known. Reactivated slide. James (1970)
(toe) investigation. suspected of being high. pressures (?)

2.5 to 2.8 (crest), Seasonal movements in winter seasons over a period of 6 years High seasonal pore
Cuffley 0 20 50 to 100 (est.) suspect slow partially broken up Phreatic surface assumed Wet weather Excavation Suspect translational along reasonably continuous discontiuity. James (1970)
3.5 m (toe) (1951 to 1957). pressures

No indication of time over which slide occurred during the winter First recorded slip in 1947. Slopes trimmed back to 4H:1V. Failed
0.7 to 0.9 (crest), Phreatic surface assumed, failure in High seasonal pore
Hadley Wood 0 19 50 to 200 (est.) suspect slow minor break up of 1951/52. Possible reactivation given soft and wet nature of unk Excavation again in winter of 1951/52. Confined to brown London Clay. Original James (1970)
1.8 (toe) winter pressures (?)
material and flatness of slope. cut in 1850, widened in 1882.
Slide occurred during the winter of 1958/59. Indication that
minor break up in Phreatic surface assumed, failure in High seasonal pore Not clear if shallow translational failure or not. Possible shallow creep
Whitstable 1.5 20 1.5 (crest), 2 (toe) 50 to 200 (est.) suspect slow movement occurred over a period of at least several weeks to unk Excavation James (1970)
toe region. winter pressures (?) movements induced by freeze / thaw action.
months.
Phreatic surface assumed, failure in
High seasonal pore Whole of cut described as unstable. Definite slips occurred in 1950 /
Grange Hill 2 20 2 (crest and toe) 100 to 400 (est.) suspect slow virtually intact Failure occurred in winter of 1950/51. Suspect slow. winter (?). Shallow counterfort unk Excavation James (1970)
pressures (?) 51. Shallow counterforts, but not effective.
drains.
Only information on movement is degraded profile indicating a
100000 to 300000 Excavation (Possibly Travel angle = 21.8 degrees. Average Su on slide surface 25 kPa (for
Isle of Sheppey - A 16 20 12.5 (toe) suspect rapid Broken up total movement of 12.5 m at the toe. Suspect the initial failure Phreatic surface assumed unk Excavation James (1970)
(est) undrained failure) undrained failure).
occurred at a rapid velocity.
Only information on movement is degraded profile indicating a
100000 to 300000 Excavation (Possibly Travel angle = 24.4 degrees. Average Su on slide surface 20 kPa (for
Isle of Sheppey - B 11.5 20 9 (toe) suspect rapid Broken up total movement of 9 m at the toe. Suspect the initial failure Phreatic surface assumed unk Excavation James (1970)
(est) undrained failure) undrained failure).
occurred at a rapid velocity.
Prior to excavation groundwater
100000 to 500000 Suspect high Failure occurred as a series of three retrogressive slips over a 24 no rainfall during No section showing post-failure profile. Likely very rapid, undrained Skempton and La Rochelle
Bradwell 17.5 20 5 to 15 (unk) suspect rapid levels 0.9 ro 1.2 m depth. Drained Excavation Excavation
(est.) degree of break up hour period. excavation failure of steep cut in fissured clay. (1965)
analysis using assumed ru value.

Nov 1948 - heaving of track below retaining wall and crack in Phreatic surface from water levels in High seasonal pore Heaving of railway track in front of wall and crack in road pavement at
Wood Green 5.5 20 0.9 (toe) 100 to 200 (est.) slow Intact unk Excavation Henkel (1957)
pavement above cut. Top of wall had moved forward ~ 0.9 m. standpipes after failure. pressures (?) crest of slope.
Section shows 2.05 m crest movement and 1.5 m movement at Assumed phreatic surface. Skempton (1977)
2.0 to 2.2 m (crest), High seasonal pore Back analysis of failed slope indicates further movements at about
Sudbury Hill 1 20 200 to 500 (est.) slow Intact toe. Further slip movements in succeeding winters from 1949 to Piezometers not installed until 7 unk Excavation Skempton (1964)
1.6 m (toe) pressures (?) residual strength.
1956. Suspect slow. years after failure. Lo and Lee (1973).
Excavation (Possibly
Movement occurred over a period of 4 hours. Described as "in Failed in winter. Possibly triggered Type 1 cut. Failure moved large distance. Travel angle = 14.6
60000 to 180000 undrained failure) and Skempton (1977)
New Cross na na 10 (crest) to 31 (toe) rapid broken up motion from top to bottom". Continued failures throughout winter by high groundwater table following heavy seasonal rain Excavation degrees. Basal surface of rupture described as "glassy like surface".
(avg.) high seasonal pore Gregory (1844)
of 1841/42. heavy rainfall. Suspect undrained failure.
pressures (?)
Series of tension cracks observed at the top of the slope. Slide Type 1 cut. Failure moved a relatively large distance due to break up
1.2 m (crest), 0.8 m Failed in winter. Possibly triggered High seasonal pore Skempton (1964)
Northolt 4 20 200 to 1200 (est) Slow to moderate Virtually intact mass did not break up at the toe and travel down the 22 degree unk Excavation at toe on 22 degree slope. Travel angle = 20.6 degrees. Material
(toe). by high groundwater table. pressures (?) Henkel (1957).
cut slope. within failure bowl remained intact.
Relatively rapid movement, slid forward 6 m in less than 0.5
Failed in winter. Possibly triggered Large failure for London Clay. Moved large distance in short time (6 m
Virtually intact, hour. Bulging in front of retaining wall. Only slight movement High seasonal pore Skempton (1977)
Wembley Hill 9 20 6m 290000 rapid by high groundwater table. Phreatic unk Excavation in 0.5 hrs). Suspect brittleness of slide associated with failure through
bulging at toe. after main slide. Loud report as cracks developed in retaining pressures (?) Anon (1918)
surface assumed. the retaining wall on the margins.
wall.

In late Sept 1954 cracks observed in road and footpath above


Failed in winter. Possibly triggered
More than 0.4 m (~ cut. Further evidence of movement in mid November prior to High seasonal pore Watson (1956)
Uxbridge - 1954 6.2 20 140 slow Intact by high groundwater table. Phreatic unk Excavation History of failures. 1937 failure was in the vicinity of this failure.
0.6 to 0.8 m ?) major movement from 23 Nov 1954. Movement occurred over pressures (?) Henkel (1956).
surface from piezometers.
several days whilst remedial measures in progress.

Section indicates that failure occurred once excavation had reached


Failure occurred during construction works for new retaining wall. Failed during excavatrion for new Excavation for new Excavation for new
Uxbridge - 1937 na na 5m 240000 to 500000 rapid partially broken up unk shear band associated with original excavation, i.e., strain weakened Watson (1956)
Temporary support of excavation failed resulting in rapid failure. retaining wall retaining wall retaining wall
zone.
Appendix D Page D31

Table D2.4: Case studies of failures in cut slopes in weathered Upper Lias clay
Age of
Slope Geometry Material Properties Slide Details
Date of Slope at Fissured" or Slope
Name Slope Height, Cut Slope % of Height Plastic Limit Undrained State of
Failure Failure "Brecciated" o Comments Liq Limit (%) Slide Classification Failure
(years) Type H s (m) ( ) Retained (%) Strength Sliding
Profile Geometry
Original cut in 1832 to 1835. Slope Compound - significant mid-
stiff to hard (surface
Stowehill - A 1901 66 Cut fissured 10.2 24.8 - above cut = 5 degrees. Cut-off 58 27 slope translational first time Type 2
of rupture - soft)
trench at crest of cut. component

Original cut in 1832 to 1835. Slope stiff to hard (surface


Stowehill - B 1957 122 Cut fissured 8.2 23 - 58 27 Compound first time Type 2
above cut = 5 degrees. of rupture - soft)

stiff to very stiff


1957, Original cut in 1896. Flat slope
Culworth - A 61 Cut fissured 8 26.5 - 65 24 (surface of rupture - Compound first time Type 2
November above cut.
firm)
stiff to very stiff
1957, Original cut in 1896. Flat slope
Culworth - B 61 Cut fissured 8 26.5 - 65 24 (surface of rupture - Rotational first time Type 2
November above cut.
firm)
Original cut in 1835. Slope above Compound - significant
Heyford 1961, January 125 Cut fissured 9.1 24 - 57 29 unk first time Type 2
cut = 8.5 degrees. translational component
Original cut in 1908. Low height
retaining wall at toe of slope (not
Ardley 1960, August 52 Cut fissured 15 22 24% involved in the failure), 6.5 degree 59 28 unk Rotational first time Type 1
slope above cut. Capped by sand
layer.

Original cut in 1877. Capped by


Hunsbury Hill Tunnel
1920 44 Cut fissured 16.4 24.5 - layer of sand. Slope above at 4.5 60 28 stiff to very stiff Rotational first time Type 2
-1920
degrees.

Original cut in 1877. Capped by stiff to very stiff, soft


Hunsbury Hill Tunnel reactivation of
1954 77 Cut fissured 16.4 24.5 - layer of sand. Slope above at 4.5 60 28 on surface of Rotational / Compound Type 2
- 1954 1920 slide
degrees. rupture.

Cut widened in 1939/40, 16 years


Charwelton 19-Apr-1955 40 Cut fissured 9.6 27.2 - prior to slip. slope above cut 59 26 unk Rotational / Compound first time Type 2
approx. 6.5 degrees.

Original cut in 1896. Capped by


Seaton 1963 67 Cut brecciated 9.3 20.8 - layer of relatively permeable head. 58 29 unk Compound first time Type 2
Slope above cut = 4 degrees.

Original cut in 1855. Partly benched


Compound - sigificant mid-
Wellingborough slope from section provided.
1961, May 105 Cut brecciated 10 31 - 61 30 unk slope translational first time Type 2
Station Capped by layer of sand, slope
component
above = 4 degrees.
Original cut in 1959. Original cut at
1.5H to 1V but failed during
1964-65, construction, then cut back to 3H to
Wothorpe - B 5 Cut brecciated 6.1 18.5 - 68 27 stiff to very stiff Translational first time Type 2
winter 1V. Capped by layer of solifluction
material. Slope above cut = 4
degrees.

Original cut in 1878. Capped by


Barrowden 1961 83 Cut brecciated 4.3 28.8 - layer of relatively permeable head. 62 31 unk Rotational first time Type 2
Slope above cut = 3.5 degrees.
Appendix D Page D32

Table D2.4: Case studies of failures in cut slopes in weathered Upper Lias (Sheet 2 of 3)
At And Post-Failure
Slide Dimensions Limit Equilibrium Analysis
Name Post-Failure Deformation Behaviour
(from James,1970)
Width Height, Depth, D Slope Volume Peak Velocity Velocity Degree of
o c' (kPa) phi ' (deg) Movement (m) Comment
(m) H f (m) (m) ( ) (cu.m.) (mm/day) Category Break-up
2 (crest), 0.5 to 1 Suspect slow movement given age at failure. Possible
Stowehill - A unk 10.2 3.9 24.8 unk 2.5 25 unk unk minor break up
(toe) series of retrogressive slips.

Stowehill - B unk 7.6 3.8 23 unk 0.5 25 0.6 (crest), 1.4 (toe) unk unk minor break up Suspect slow movement given age at failure.

2 (crest), 1.5 to 2 suspect slow movement given age of cut and low
Culworth - A unk 7.4 4.2 26.5 unk 1.9 25 unk unk virtually intact
(toe) cohesion.

2 (crest), 0.5 to 1.6 suspect slow movement given age of cut and low
Culworth - B unk 9 4.8 26.5 unk 0.7 25 unk unk virtually intact
(toe) cohesion.

2.0 (crest), 1.5 to


Heyford unk 9.7 3.2 24 unk 2.5 25 unk unk virtually intact suspect slow movement given age of cut.
2.3 (toe)

Ardley unk 4.9 3.1 22 unk 0 25 0.5 (crest), 0.8 (toe) unk unk virtually intact suspect slow due to limited deformation.

Hunsbury Hill
unk 15.8 6.1 24.5 unk 4 25 na unk unk unk Limited information available
Tunnel -1920

Numerous failures over period from 1900 to 1954.


Hunsbury Hill 7.6 (crest), 0.8 to 1 suspect slow to Crest movement totalled 7.6 m. Suspect cutting back
unk 15.8 6.1 24.5 unk 0 18 unk minor break up.
Tunnel - 1954 (toe) moderate at toe for removal of slide debris affected stability of
slope for failure in 1954.

Slide active for more than 1 week, therefore suspect


Charwelton 46 9.6 5.4 27.2 2500 to 3000 2.4 25 1.4 (crest), 1.3 (toe) 200 to 400 suspect slow virtually intact
slow.

0.3 to 2.0 (crest),


Seaton unk 9.8 3 20.8 unk 2.5 23 unk unk virtually intact Limited information.
0.7 to 2.2 (toe)

Limited information available. Initial failure in 1961.


Wellingboroug
unk 11 4 31 unk 0 23 unk unk unk unk Reactication of movement during 1964 remedial
h Station
works.

3.5 (crest), 1.5 to 2 20 to 100 suspect slow (initial Movement initiated in 1964-65 winter and continued to
Wothorpe - B unk 6.1 1.6 18.5 unk 1 23 virtually intact
(toe) (reactivation) slide) move thereafter. 1.5 m after 3 years.

0.55 (crest), 0.5


Barrowden unk 4.95 4.7 28.8 unk 0 23 unk suspect slow virtually intact Suspect slow due to limited amount of movement.
(toe)
Appendix D Page D33

Table D2.4: Case studies of failures in cut slopes in weathered Upper Lias (Sheet 3of 3)
Hydrology
References
Name Hydro-geological Cause/s of Sliding Trigger of Slide Comments
Rainfall / Snowmelt (Author/Year)
Features

No information, phreatic surface High seasonal pore Original cut in 1832 to 1835. First failure in 1901 followed by Chandler (1972)
Stowehill - A unk Excavation
assumed for stability analysis. pressures (?) subsequent larger failure (Stowehill - B) in 1957. Chandler (1974)

No information, phreatic surface


High seasonal pore Original cut in 1832 to 1835. This failure partly encompasses Chandler (1972)
Stowehill - B assumed for stability analysis. unk Excavation
pressures (?) 1901 failure, but is more extensive. Chandler (1974)
Counterfort drains on slope.

Previous summer had been very


High seasonal pore Original cut in 1896. Failures A and B were only 80 m apart. Chandler (1972)
Culworth - A wet, possibly high piezometric very wet summer Excavation
pressures (?) High water content observed on surface of rupture. Chandler (1974)
surface.
Previous summer had been very
High seasonal pore Original cut in 1896. Failures A and B were only 80 m apart. Chandler (1972)
Culworth - B wet, possibly high piezometric very wet summer Excavation
pressures (?) High water content observed on surface of rupture. Chandler (1974)
surface.
Limited information on water High seasonal pore
Heyford Very high Dec 1960 rainfall Excavation Original cut in 1835. Possible slip prior to 1961. Chandler (1974)
levels. pressures (?)

Piezometric levels from Heavy storms in August High seasonal pore Original cut in 1908.Type 1 cut slope failure in slope above
Ardley Excavation Chandler (1974)
piezometers installed in 1969. 1960 (76 mm on 7/8/60) pressures (?) retaining wall. Capped by sand layer.

no information. Piezometric Original cut in 1877. History of slope failures with failures in
Hunsbury Hill Tunnel - High seasonal pore
levels for stability analysis unk Excavation 1900, 1920, 1921, 1928, 1935, 1945 and 1954. Slope then Chandler (1974)
1920 pressures (?)
assumed. trimmed back.

no information. Piezometric
Hunsbury Hill Tunnel - High seasonal pore Original cut in 1877. History of slope failures. This failure
levels for stability analysis unk Excavation Chandler (1974)
1954 pressures (?) reactivation at residual strength.
assumed.

Cut widened in 1939/40, 16 years prior to slip. Slip active for


Phreatic surface for stability High seasonal pore
Charwelton unk Excavation at least one week. Possibly deeper than indicated in figure James (1970)
analysis assumed. pressures (?)
because movement of railway tracks observed.

Piezometers installed in 1969,


High seasonal pore
Seaton maximum levels used for unk Excavation Original cut in 1896.Translational failure in brecciated clay. Chandler (1974)
pressures (?)
stability analysis.

no information. Piezometric
no exceptional rainfallprior to High seasonal pore Original cut in 1855. Reactivation in 1964 after remedial
Wellingborough Station levels for stability analysis Excavation Chandler (1974)
slip pressures (?) works.
assumed.

Original cut in 1959. Failures occurred during construction in


Piezometric levels from High seasonal pore Chandler (1972)
Wothorpe - B unk Excavation slopes at 1.5H to 1V. One of these failures continued to
piezometers. pressures (?) Chandler (1974)
degrade even after trimming back to 3H to 1V.

Piezometers installed in 1969,


High seasonal pore Original cut in 1878. Large seasonal variation in piezometric
Barrowden maximum levels used for unk Excavation Chandler (1974)
pressures (?) levels near to ground surface (virtually no change at depth).
stability analysis.
APPENDIX E

Case Studies of Concrete Face


Rockfill Dams
Appendix E Page E1

1.0 CASE STUDY INFORMATION ON CONCRETE FACE


ROCKFILL DAMS
Information on the individual case studies used for the analysis of rockfill deformation
behaviour in embankment dams is presented in Table E1.1. The data set comprises
thirty-five concrete faced rockfill dams and one central core earth and rockfill dam (El
Infiernillo dam).
An explanation of some of the terminology used in Table E1.1 is as follows:
• General terms or symbols used and their meaning are:
− “-” or "na" = not required, not appropriate or not known
− “( )” in strength class, Cu, dry density and void ratio columns indicates the
value has been estimated.
− H to V refers to horizontal to vertical slope angle; i.e. 2H to 1V.
− SMP = surface monitoring point.
− HSG = hydro-static settlement gauge
− IN = inclinometer
− EOC = end of construction
− FF = first filling
• The system for classification of the various rockfill zones used for the main rockfill
is described in Section 6.2.4 of Chapter 6. The zoning classification used by the
author/s of the referenced paper is given separately.
• Under the section construction:
− Height, H = maximum embankment height
− Length, L = crest length
− Slab face area = the area of the upstream face in the plane measured normal to
the upstream face.
• In the design section:
− Main rockfill refers to the rockfill forming the main body of the embankment;
i.e. Zone 3 for dumped rockfill, Zones 3A and 3B for compacted rockfill
(refer Section 6.2.4 of Chapter 6 for details).
− Facing details refers to the embankment zone between the facing slab and the
main rockfill (refer to Section 6.2.4 of Chapter 6).
• In the section on rockfill material parameters and properties:
Appendix E Page E2

− Terms FR and SW, used in rockfill source column, refer to the rock
weathering classification (fresh and slightly weathered respectively) used by
the referenced author/s.
− Classification of the unconfined compressive strength (UCS) of the intact
rock used in the rockfill is in accordance with Australian Standard AS 1726-
1993 (refer Table 1.2 in Section 1.3.3 of Chapter 1).
− Particle size distribution parameters have been determined from the average
grading curve for the rockfill where available or as reported by the referenced
author/s.
− Cu = Grading uniformity coefficient of the rockfill particle size distribution
(= D60/D10).
− Cc = Grading curvature coefficient of the rockfill particle size distribution

(= ( D30 ) ( D60 ⋅ D10 ) ).


2

− D10, D30, etc., refer to the particle size (from the average grading curve)
related to 10, 30, etc. percent passing.
− The dry density, void ratio and porosity are average values reported by the
referenced author/s or calculated from the available data.
− The classification terms used to describe the “level of compaction” are
defined in Section 1.3.3 of Chapter 1. “Good” refers to well-compacted,
“poor” refers to dumped and well sluiced or poorly compacted, and “very
poor” refers to dry dumped or dumped and poorly sluiced rockfill.
− 4p 10t SDVR = 4 passes of a 10 tonne (dead-weight) smooth drum vibrating
roller.
• Definitions of and methods for calculation of the rockfill secant moduli during
construction (Erc) and the modulus on first filling (E rf) are discussed in Chapter 6
(Sections 6.2.3 and 6.4.2). The rockfill secant modulus during construction (Erc)
has generally been determined from deformation records in the lower 60% and
central half of the embankment. In Table E1.1:
− The “average [secant modulus] over the construction period” has determined
by averaging the modulus estimates throughout the construction period. It
includes values estimated when the embankment was in the early stages of
construction right through to the end of construction; and
Appendix E Page E3

− The “average at EOC (end of construction)” only considers those secant


moduli estimates calculated at the end of the main rockfill construction, the
calculation for which takes into consideration the effect of valley shape on the
vertical stress. The “average” is a representative value of the modulus
obtained by averaging where more than one estimate has been made (i.e. from
or between different HSGs).
• Leakage. The time period of the long-term leakage rate is generally given in years
after the end of first filling. For several cases, such as Dix River, it is given in terms
of years after the end of main rockfill construction because the period of first filling
was not known. For several embankments where first filling was not completed in
the reported period and where a significant reduction in leakage rate occurred, such
as due to repair, a long-term leakage rate is given (e.g. Khao Laem).
• Deformation monitoring:
− The settlements during construction are generally based on the monitoring
records from hydrostatic settlement gauges (HSG) installed as construction
proceeded;
− Post construction crest settlements are from after the end of main rockfill
construction. “0.5 to 5 years” indicates the period of monitoring after the end
of main rockfill construction;
− Post construction crest settlements are given in millimetres and as a
percentage of the embankment height;
− For the deformation normal to the face slab, the references to percent of dam
height or distance below the crest (“bc” = vertical distance below the crest)
refer to the location on the face slab representative of the location of the
deformation measurement.
− The long-term deformation rates are in units of crest settlement (as a
percentage of the embankment height) per log cycle of time, where time is
measured in years after the end of main rockfill construction.
Appendix E Page E4

Table E1.1: Case studies information on concrete face rockfill dams


GENERAL DETAILS CONSTRUCTION DESIGN
Construction Timing Dimensions Facing Details (Support Zone for Face Slab)
Name Location /
Geology
Foundation
Main Rockfill (Comments) Thickness References
Owner Year Time Height, Length, Slab Face Upstream Downstream (Soil/Rock) Material Type /
L/H (normal to face Comments
Completed (years) H (m) L (m) Area (m2) Slope Slope Compaction
slab), m
Aguamilpa (Zone 3A) Macedo et al (2000), Macedo (1999)
Rock - gravels in Zone 3A upstream half. Zone 3T dowmstream central River gravels (sandy
Tertiary volcanics - Montanez et al (1993), Mori (1999)
Aguamilpa (Zone 3T) CFE, Mexico 1993 3? 185.5 475 2.6 130000 1.5 to 1 1.4 to 1 part of river 1/3. Zone 3B - downstream outer 2/3. Zone 3B coaser - gravel), rounded, well -
ignimbrites Gonzales & Mena (1997)
section than 3T. compacted.
Marulanda & Pinto (2000)
Aguamilpa (Zone 3B)

metamorphic sediments, Gravels in river Zone 3A upstream half of main rockfill, Zone 3B finer well graded rockfill,
-
Materon (1985a), Regalado et al (1982)
Alto Anchicaya Columbia schists & chert
1974 1.5 140 260 1.9 31000 1.4 to 1 1.4 to 1 downstream half of main rockfill.
1.75 to 7 well compacted Amaya & Marulanda (2000)
section
Zone 3A - 90% of main rockfill, Zone 3B in lower part of
finer well graded rockfill, higher densities than HEC (1991a), Bowling (1981-82)
Bastyan Hydro Tasmania Rhyolite 1983 1.5 75 430 5.7 19000 1.3 to 1 1.3 to 1 Rock outer downstream shoulder. Slight influence from 3.7 well compacted main rockfill Knoop and Lack (1985)
arching in lower gully.
Forza & Hancock (1995a), Liggins (1971),
Quartzite & quartzite Zone 3A - 75% of main rockfill, Zone 3B in outer finer well graded rockfill, Fitzpatrick et al (1973), Wilkins et al (1973),
Cethana Hydro Tasmania conglomerate
1971 2.5 110 213 1.9 30000 1.3 to 1 1.3 to 1 Rock downstream shoulder. Likely arching in lower gully.
3.7 well compacted
-
Fitzpatrick et al (1982), Pinto et al (1982),
Giudici et al (2000)

Zones 3B and 3C (Zone 3A) similar in rockfill type and crushed rockfill, finer than Higher densities than Wu & Cao (1993)
Chengbing China Tuff lava 1989 2 to 2.5 74.6 325 4.4 15800 1.3 to 1 1.3 to 1 Rock 2.4
compaction 3A and well compacted main rockfill Wu et al (2000c)

Laminated gunite facing damaged early in embankment Lean concrete facing


Cogswell LACFCD Granitic Gneiss ? 1935 3.3 85.3 176.8 2.1 - 1.3 to 1 1.5 to 1 Rock life. Replaced in 1935 with timber facing and in 1947 1.8 to 5.5 derrick placed rockfill placed on derrick Baumann (1939, 1958, 1964)
with reinforced concrete facing. rockfill

Pacific Gas &


derrick placed rockfill, voids Cooke (1958)
Courtright Electricity Co. granite 1959 2 97 274 2.8 20440 1.15 to 1 1.4 to 1 Rock - 2.4 to 3.5 chinked
-
Regan (1997)
San Francisco
tunnel spoil, rockfill Davies (1993), Quinlan (1993)
sandstone, phyllite, Facing rockfill (Zones 2A and 2B) likely to have a lower rockfill used, well
Crotty Hydro Tasmania 1991 2 83 240 2.9 14500 1.3 to 1 1.5 to 1 Rock 7.3 from sedimentary Li et al (1993), Cribben (1990)
conglomerate modulus than the gravels used in the main fill (Zone 3A). compacted
geology units. Giudici et al (2000)
Kentucky derrick placed rockfill, 2.7 to
Suspect rockfill likely to be relatively coarse. Some Schmidt (1958)
Dix River Utilities limestone & shale 1925 2 84 311 3.7 - 1.1 to 1 1.4 to 1 Rock 1.4 to 4.1 9 tonne rocks used, voids -
rockfill placed directly from fired shots in abutments. Howson (1939)
Company chinked

El-Infiernillo (Zone 3B) Central core earth and rockfill embankment. Included to
CFE, Mexico silicified conglomerate 1963 1.3 148 344 2.3 - 1.75 to 1 1.75 to 1 Soil/Rock give an estimate of moduli during construction for dry na na na Marsal & Ramirez de Arellano (1967)
El-Infiernillo (Zone 3C) dumped rockfill.

Foz Do Areia (Zone 3A) Complex design. Zone 3A on roughly upstream 1/3, finer graded sound basalt Pinto et al (1985a), Materon (1985b)
Basalt & Basaltic
Brazil 1979 2.5 160 828 5.2 139000 1.4 to 1 1.4 to 1 Rock Zones 3B (1C and 1D) in downstream 2/3 of main 2.3 to 7.6 (very high strength), well Pinto et al (1982), Pinto et al (1993)
Breccia rockfill. compacted Sobrinho et al (2000), Cooke (1999)
Foz Do Areia (Zone 3B)

processed dirty gravels,


interbedded shale and Clean gravels placed as chimney filter and in central
finer graded, well
slightly higher density Amaya & Marulanda (1985)
Golillas Columbia 1978 1.75 125 107 0.9 17000 1.6 to 1 1.6 to 1 Rock section of main fill.
4.0 than main rockfill Amaya & Marulanda (2000)
sandstone compacted

Zone 3A - upstream 1/3; Zone 3B - downstream 2/3.


Ita (Zone 3A) Sobrinho et al (2000)
Dumped rockfill placed in river section. Zone 3A crushed dense basalt, fine high densities than
Brazil Basalt 1999 3 125 880 7.0 110000 1.3 to 1 1.3 to 1 Rock 3.5 to 5 Sobrinho et al (1999)
incorporates dense basalt zone closest to upstream face sized, well compacted main rockfill
Ita (Zone 3B) Silveira & Sardinha (1999)
and zone of lesser quality behind this.

Finer graded H to VH Good et al (1985), Good (1976, 1981)


SA Water, South Rock, some Zone 3A is main body of rockfill. Zone 3B is a basal Zone 2 placed in 915
Kangaroo Creek schist and gneiss 1969 1.75 60 178 3.0 - 1.3 to 1 1.4 to 1 8.0 strength granitic gniess, E&WSD (1981), SKM (1995)
Australia gravels layer of very high strength rockfill for drainage purposes.
well compacted
mm layers.
SA Water (1995)

Khao Laem (Zone 3A) limestone, shale, Zone 3A - 75% of main rockfill in section, Zone 3B in finer graded rockfill, 0.5 m
Watakeekul et al (1985)
Thailand 1984 2 130 1000 7.7 - 1.4 to 1 1.55 to 1 Rock, karstic 4.1 D50 = 24 to 40 mm Mahasandana & Mahatharadol (1985)
sandstone & siltstone outer downstream shoulder (25% of main rockfill). layers, well compacted
Aphaiphuminart et al (1988)
Khao Laem (Zone 3B)
Gosschalk & Kulasinghe (1985, 1987)
Limited information on rockfill properties. HSGs in higher finer graded rockfill, well
Kotmale Sri Lanka foliated gneiss 1984 3 90 560 6.2 - 1.4 to 1 1.45 to 1 Rock 3.5 - Casinader (1987)
elevations indicate higher modulus. compacted
Kulasinghe & Tandon (1993)
High strength (UCS = 36 to Better quality rockfill
SA Water, South interbedded quartzite, No water available to add to rockfill for construction to 24 Good et al (1985), Good (1981)
Little Para 1977 2.5 53 225 4.2 - 1.3 to 1 1.4 to 1 Rock m height.
8.5 45 MPa) quartzite rockfill, used for facing
Boucaut & Beal (1979)
Australia slaty dolomite & slate well compacted layers.

Lower Bear No. 1 Pacific Gas & granite 1952 1.5 75 293 3.9 17650 1.3 to 1 1.35 to 1 Rock Coarse rockfill 3 to 6.4 derrick placed rockfill, 1.15
3
Electricity Co. to 2.7 m volume, voids - Steele & Cooke (1958)
Lower Bear No. 2 San Francisco granite 1952 <1.5 46 264 5.7 8400 1 to 1 1.3 to 1 Rock Coarse rockfill 3 to 5.2 chinked
Appendix E Page E5

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 2 of 8)
Material Parameters / Properties of Rockfill Material Parameters / Properties of Rockfill
Rockfill Zone Intact Rock Strength Particle Size Distribution Parameters (from average grading curve) Particle Size Distribution Parameters (from average grading curve)
Name Rockfill Particle
Wet/Dry % finer % finer
n Authors Source Shape Strength Absorption D 10 D 30 D50 D 60 D 80 D90 Dmax % finer Clean / D 10 D 30 D50 D 60 D 80 D90 Dmax % finer Clean /
Class UCS (MPa) Strength Cu Cc 75 Cu Cc 75
Classn Classification (%) (mm) (mm) (mm) (mm) (mm) (mm) (mm) 19 mm Dirty (mm) (mm) (mm) (mm) (mm) (mm) (mm) 19 mm Dirty
(%) micron micron
dredged alluvium rounded 34
clean to 34
clean to
Aguamilpa (Zone 3A) 3A 3B (Very High) - - - 85 5.8 0.75 16.4 45 62 124 230 600 2 85 5.8 0.75 16.4 45 62 124 230 600 2
dirty dirty
Aguamilpa (Zone 3T) 3T T ignimbrite angular Very High 180 - - 30 2.4 2.9 24.5 61 88 204 290 500 2 28 clean 30 2.4 2.9 24.5 61 88 204 290 500 2 28 clean

Aguamilpa (Zone 3B) 3B 3C ignimbrite angular Very High 180 - - 22 1.4 9 - 150 - - - 700 - - clean 22 1.4 9 - 150 - - - 700 - - clean

Alto Anchicaya 3A D hornfels angular (Very High) - - - 18 2.5 5.5 31 71 100 205 280 600 <2 22 clean 18 2.5 5.5 31 71 100 205 280 600 <2 22 clean

Rhyolite (SW to
Bastyan 3A 3A FR)
angular (Very High) - - - 42 3.3 2.4 28 69 100 190 300 600 3.3 25 clean (?) 42 3.3 2.4 28 69 100 190 300 600 3.3 25 clean (?)

Cethana 3A 3A Quartzite, quarried angular (Very High) - - - 23 2.0 4.8 32 75 108 250 440 900 2 21 clean 23 2.0 4.8 32 75 108 250 440 900 2 21 clean

Chengbing 3A 3B/3C Tuff lava, quarried angular Very High 80 - - 10.4 - - - - - - - 1000 - - clean (?) 10.4 - - - - - - - 1000 - - clean (?)

Cogswell 3A 3A Granitic Gneiss angular High 45 - - 7 2.3 100 - 650 - - - 1300 3 5 clean 7 2.3 100 - 650 - - - 1300 3 5 clean

550 to 550 to
Courtright 3A 3A granite, blasted angular (Very High) - - - (<7) - - - - - - 1750 - - clean ? (<7) - - - - - - 1750 - - clean ?
1500 1500

Gravels
Crotty 3A 3A rounded (Very High) - - - 70 5.6 0.34 7.3 20 28 60 86 200 3.5 48 dirty 70 5.6 0.34 7.3 20 28 60 86 200 3.5 48 dirty
(Pleitocene)

Limestone & Shale


Dix River 3A 3A angular ? (Very High) - - - low - - - - - - - - - - - low - - - - - - - - - - -
(?), spillway excav

diorite & silicified


El-Infiernillo (Zone 3B) 3B 3B conglomerate
angular Very High 125 - - 13 1.6 6.5 - 62 - - - 600 1 22 clean ? 13 1.6 6.5 - 62 - - - 600 1 22 clean ?

diorite & silicified


El-Infiernillo (Zone 3C) 3C 3C conglomerate
angular Very High Diorite = 125 - - <13 1.6 ? >7 - - - - - >600 - - clean ? <13 1.6 ? >7 - - - - - >600 - - clean ?

basalt (max 25%


angular
High to Very Basalt = 235 10 10
Foz Do Areia (Zone 3A) 3A 1B basaltic breccia)
- 1.0 6 2.1 20 60 105 145 300 440 600 <1 clean 6 2.1 20 60 105 145 300 440 600 <1 clean
High Breccia = 37
mix basalt & High to Very Basalt = 235
Foz Do Areia (Zone 3B) 3B 1C & 1D basaltic breccia
angular - 1.0 14.2 2.2 - - - - - - - - - clean ? 14.2 2.2 - - - - - - - - - clean ?
High Breccia = 37

gravels,
Golillas 3A 2 rounded (Very High) - - - 125 2.5 0.65 11.5 31 80 185 225 350 6 40 dirty 125 2.5 0.65 11.5 31 80 185 225 350 6 40 dirty
unprocessed

Ita (Zone 3A) 3A E1/E3' basalt (dense) angular (Very High) - - - 11 1.6 13.4 57 113 152 275 425 700 1 12 clean 11 1.6 13.4 57 113 152 275 425 700 1 12 clean

Breccia and Basalt angular


Basalt - VH 15 15
Ita (Zone 3B) 3B E3 - - - 13.3 1.6 10 46 100 133 250 385 750 1 clean 13.3 1.6 10 46 100 133 250 385 750 1 clean
Breccia - H (?)

angular to
Kangaroo Creek 3A 3 schist Medium to High 25 63 3.0 310 6.4 0.14 6.2 26 44 105 180 600 8.5 44 dirty 310 6.4 0.14 6.2 26 44 105 180 600 8.5 44 dirty
subangular

Khao Laem (Zone 3A) 3A 3A limestone, quarried angular Very High < 190 100 low - - - - - - - - 900 - - - - - - - - - - - 900 - - -

Khao Laem (Zone 3B) 3B 3B limestone, quarried angular Very High < 190 100 low - - - - - - - - 1500 - - - - - - - - - - - 1500 - - -

charnockitic /
angular
(High to Very - -
Kotmale 3A 3A gneissic
- - - - - >2 - - - - - 700 - - - - >2 - - - - - 700 - -
High)

angular, close to dirty (soft dirty (soft


Little Para 3A 3A dolomitic siltstone
elongated ?
Medium 8 to 14 - 120 4.0 0.6 13 40 70 255 390 1000 5 35
rock)
120 4.0 0.6 13 40 70 255 390 1000 5 35
rock)
100%

Lower Bear No. 1 3A 3A granodiorite angular Very High 100 to 140 - -


700 to 2000 to 700 to 2000 to
8 to 10 - >100 - - - - low low clean 8 to 10 - >100 - - - - low low clean
800 3000 800 3000
Lower Bear No. 2 3A 3A granodiorite angular Very High 100 to 140 - -
Appendix E Page E6

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 3 of 8)
SECANT MODULI VALUES HYDROGEOLOGY LEAKAGE
Modulus During, Erc (MPa) Modulus on First Filling, E rf Time for First Filling Rainfall,
Name Maximum
(from end of main average Long-term
Average at EOC Reservoir Level / Comments Rate Comment
Average over rockfill annual (l/sec)
(adjusted for Erf (MPa) Comments (l/sec)
construction period construction) (mm)
valley shape)
Aguamilpa (Zone 3A) 305 (250 to 330) 305
Reduced modulus where Zone 3C has an influence on 200, at first 150 to 170 (4 Large increase in leakage rate when reservoir raised above
Aguamilpa (Zone 3T) 104 104 770 the deformation of the face slab.
0.48 to 1.83 years - - certain level (due to cracking in upper part of face slab)
filling years post ff)
Aguamilpa (Zone 3B) 36 (25 to 45) 36

375 (@ 30 to High rate of leakage on first filling, through the perimeter joint on
Alto Anchicaya 138 (100 to 170) 110 - 0.18 to 0.2 (Oct 74) years. - - 1800 - the right abutment and open joints on left abutment. Repaired.
40% height)

5 (8 years post Approximately 10 l/sec on first filling. Gradual reduction with time
Bastyan 130 (120 to 140) 102 290 Maximum face deflection at about mid height. 0.86 to 1.05 years. Steady 2100 10
ff) to base flow of 5 l/sec.

160 (120 to 210) in lower 25%


5 to 10 (4 years
Cethana 105 (85 to 120) in mid to lower 103 300 - 0.26 to 0.48 years (1971). Steady 2030 65 to 70 -
25%
post ff)

First filling started before embankment


-0.5 to 4 years (12/88 to Max leakage during construction when reservoir level rose above
Chengbing 43 43 110 Approx 2.5 times E during completed. Fluctuating, drawdowns up to - 75 -
1993). level of face slab. After completion leakage = 60 l/sec.
50 m.
Started filling 2.6 years
Cogswell - - 44 From measured face slab deformation. after EOC. Filled in 71 - - 3600 - Maximum in 1938 on first filling. Large crack in floater slab.
hours in March 1938.

Lower perimetric joints moved soon after completion affecting


Courtright - - - - 0 to 9 years (from 1959). Fluctuating - 1275 - waterstops. Joints on abutments opened (~ 1968) and upper
perimetric joint opened tearing waterstops in 1989.

Relatively low modulus (compared with modulus during


construction) may be associated with lower modulus of 0.87 to 2.0 years. Steady
30 to 35 (2 years Peak leakage rate of 45 l/sec. Thought to be seepage through
Crotty 375 (113 to 636) 360 470 2900 45 the foundation on the abutments where the grouting failed.
facing rockfill.
post ff)
2700 (1937, Leakage near intake tower, through cracks and opened joints in
2000 (30 years
Dix River - - - - Started in 1925 - - 12 yrs after slabs and through the foundation. Various methods to try and
post EOC) treat problem.
EOC)

El-Infiernillo (Zone 3B) 39 (27 to 48) 35.5


- - 0.33 yrs (June to Oct 1964) Fluctuating - - - -
El-Infiernillo (Zone 3C) 22 (17 to 27) 20

Foz Do Areia (Zone 3A) 47 (38 to 56) 47 Relatively large deformations in toe region compared with
other CFRDs. Therefore lower moduli toward toe region
0.5 to 0.9 years (Apr to
Relatively steady, typical 5 m fluctuation
70 (5 years post Maximum leakage on first filling. Gradual reduction with time.
80 (65 to 92) Aug 80).
- 240
(15% height above toe)
ff)
Foz Do Areia (Zone 3B) 32 (29 to 38) 32

155 (145 to 165)


Estimated by hydrostatic settlement cells for section 4 to 5.17 years (June 82 to Fluctuating. First filling interupted due to 270 to 500 (15 Leakage through the foundation during first filling and some
Golillas excludes those points thought to 107 250 - 1080
be affected by arching
below reservoir level. Aug 83). excessive seepage. years post EOC) segregation of Zone 1 noted. Repaired and leakage reduced.

Ita (Zone 3A) 48 48 Reached 1700 l/sec 4 months after first filling, due to cracking in
Relatively low because incorporates low modulus
87 (83 to 91) 0.7 to 0.9 years. - - 1700 - slab. Reduced to 380 l/sec by dumping sand and silty sand on
dumped rockfill in river gully (E during = 17 MPa).
Ita (Zone 3B) 24 (14 to 46) 24 face.

0 to 1.95 yrs (Sept 69 to Fluctuating, seasonal drawdown of up to 42 On initial filling seepage heard through within the embankment.
Kangaroo Creek - - 140 Reported (no data to confirm)
Aug 71). m
600 to 800 11 2.5 Possible seepage along horizontal layers in weak rockfill.

Khao Laem (Zone 3A) 59 (43 to 79) 43 High reading possibly affected by arching in deepest 1.1 (after repair Feb 1986 (3 months post first filling) leakage increased from 9 to
0.6 to 1.9 years (June 84 to
130 to 240 section of embankment, low reading possibly affected by
Nov 85).
Steady (?) - 53 of face slab 53 l/sec as a result of cracks in the face slab. Cracks repaired
Khao Laem (Zone 3B) 30 30 "plum pudding" foundation. cracks) and leakage reduced.

145 (135 to Estimated from HSGs near to upstream face.


0.54 to 1.04 years (Nov 84
Fluctuating, seasonal drawdown up to 60 m At close to end of first filling.
Kotmale 61 (47 to 87) 52 to Aug 85).
- < 10 -
155)
Maximum leakage measured just prior to first overflow in August
0.58 to 2.83 years (Aug 77
Little Para 21.5 (19.5 to 23.5) 20.5 - No data for calculation
to Nov 79).
Seasonal fluctuation of about 5 m. - 19.2 - 1981. Wet spots and seepage on downstream slope on right
abutment.
At end of 2 yrs and 42 m below crest (39% dam height). 0.07 to 0.58 years (Dec 52 Fluctuating, seasonal drawdown typically 43 (4 years post
Lower Bear No. 1 - - 21 - 113 Maximum at FSL on first filling. Gradual reduction thereafter.
Likely under-estimate to June 53). 35 m ff)
0.07 to 0.58 years (Dec 52 Fluctuating, seasonal drawdown typically Suspect relatively small in comparison to Lower Bear No. 1 and
Lower Bear No. 2 - - 40 At end of 1 year and 29 m below crest (39% dam height). - no leakage -
to June 53). 35 m other CFRDs constructed in the 1950s.
Appendix E Page E7

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 4 of 8)
MONITORING
Settlement Post Construction Crest Deformation Long Term Crest
Deformation Normal
Name During
Settlement Settlement Lateral to Face Slab
Settlement Rate
Comments
Other Comments
Construction Settlement, total Settlement, (% per log time cycle)
on FF on FF Displacement (mm)
(mm) (mm) total (%) ( t o at EOC)
(mm) (%) (mm)
Aguamilpa (Zone 3A)
Decreasing velocity of deformation with time. Face slab deformation greatest at
Aguamilpa (Zone 3T) - 307 (0.4 to 5.7 yrs) 0.185 222 0.133 - 320 from ff to 6 years 0.090 crest (due to much lower modulus of Zone 3C rockfill). Limited time for estimation Moduli during construction from literature.
of long term creep rate.
Aguamilpa (Zone 3B)

153 mm (0.11 to 10.3 160 mm to 5/77 (130 mm Decreasing rate of deformation with time. At 10 yrs < 2 mm/yr crest settlement.
Alto Anchicaya - 0.109 15 0.011 - 0.037 Crest points on top of concrete plinth.
Moduli during construction from literature.
yrs) ff)

50 mm (0.2 to 7.5 16 mm (SC3), 7 68 mm @ 34 m below


Bastyan > 219 0.066 15 0.02 0.028 Decreasing rate of deformation post first filling (also decreasing on log scale).
years, SC3) mm ff crest

137 mm (0 to 28.6 83 mm to 7/92 170 mm to 10/93 (114 ff Decreasing rate of deformation post first filling (linear on log scale). Upper 12 m of
Cethana > 560 0.124 46 0.043 0.042
yrs) (27 ff) 37m bc) rockfill placed 1 year after main body had been completed.

154 mm after 10 years Insufficient information on deformation.


In 1989, during construction, the impounded water level exceeded the top level of the
Chengbing - - - - - - - constructed face slab resulting in leakage through the embankment above the slab.
(?)

In late 1933 a very hevy rain storm resulted in crest settlements of to a maximum
Estimated at 271 mm (0 to 3 403 mm vertical settlement
Cogswell 0.317 < 271 < 0.317 - to 4/38 (first filling)
- of 3.41 m (due to collapse settlement). Artificial sluicing resulted in an additional
15.7 m. years, to 4/38) 1.71 m.

1237 mm (0 to 38 0.70 (from 5 year to 20+ Relatively steady rate of crest settlement on log scale, but only limited number of
Courtright - 1.282 844 0.875 - - No data to calculate modulus during construction or on first filling
yrs) years) data points.

55 mm (0 to 8.5 Gradual reduction in rate of deformation after first filling, relatively steady on log
Crotty > 179 0.066 16 0.019 9 mm (SC3 - ff) 46 mm @ 49 m bc 0.045 scale.
years, SC4)

1281 mm (Stn 16.9, 970 mm to 1957 0.58 to 1.44 (increasing Decreasing rate of deformation with time. At 30 yrs approximately 20 to 25
Dix River - 1.525 - - - mm/year settlement, and 15 to 17 mm/yr displacement.
No data to calculate modulus during construction or on first filling
to 1957, 32 yrs) (Stn 16.9) with time)

El-Infiernillo (Zone 3B) 1095 (1.9%) in


IN3, and 2425 Central core earth and rockfill dam. Data used for analysis of rockfill deformation during
- - - - - - na -
mm (2.2%) in construction.
El-Infiernillo (Zone 3C) IND1.

Foz Do Areia (Zone 3A) 248mm to 1984 Log scale deformation plot shows sharp reduction in rate of deformation after 3
328 mm (0 to 11 780 mm @ 49% dam years and then acceleration in rate beyond 7 years (to that similar from 1 to 3
- 0.205 73 0.046 (180 ff), d/stream 0.099
years) height (620 mm ff) years).
Foz Do Areia (Zone 3B) slope

160 mm @ 46% dam Acceleration in crest settlement and face deflection toward end of first filling.
52 mm (0.46 to 6.4 7 mm to 10/84 (2
Golillas - 0.042 20 0.016 height by 3/84 (first 0.033 Thereafter, significant reduction in rate of deformation, but limited data. Crest
years, to 10/84) yrs) SMPs on parapet. Upstream face slab deformation estimated by HSGs and SMPs.
filling)

Ita (Zone 3A)


Relatively complex cross section. Face slab cracked resulting in relatively high leakage
- - - - - - 461 mm (during ff) - Limited data.
rates.
Ita (Zone 3B)

Not sure when monitoring first started, possibly shortly after start of first filling or
50 mm to 1979 Initial monitoring possibly missed first filling to EL 220 (26 m below crest, 22 m below
Kangaroo Creek - 116 mm (0 to 26 yrs) 0.193 26 0.043 - 0.126 could have missed most of first filling entirely. Decreasing rate of movement with
FSL), but cannot be sure. Dam raised approx. 3.4 m in 1983 for additional flood storage.
(8yrs), 32 mm ff time post first filling. Crest SMPs on rockfill and face slab.

Khao Laem (Zone 3A) > 550 (of Calculation of moduli during construction takes into consideration foundation settlement.
- - - - - - - Significant settlement of foundation during construction.
Erf adjusted for filling to full supply level.
embankment)
Khao Laem (Zone 3B)

255 mm (0 to 2.46 62 mm to 10/86 98 mm to ff (estimated from 0.258 (limited data, possibly Initial reading very close to time of completion of rockfill. Last available readings
Modulus on first filling determined from deformation of HSGs adjacent to face slab.
Kotmale > 1020 0.283 96 0.107 HSG) hasn't steadied as yet) approx 12 months after first filling, rate has not settled down as yet.
yrs) (0 to 2 yrs)

152 mm (0 to 22.6 Monitoring appears to have started about June 1978, most likely when first filling Wet spots and seepage observed on downstream slope (on right abutment). Indicative of
Little Para - 0.288 22 0.042 - - 0.21 started. high permeability variation between horizontal and vertical directions.
yrs)

375 mm (0.07 to 4.05 305 mm to 11/56 (4 625 mm to 10/54 (2 yrs) @ Rapid deformation on first filling with significant reduction in rate thereafter. Crest
Lower Bear No. 1 - 0.56 < 335 < 0.50 0.103
years, to 11/56) yrs), ff 270 mm 42 m bc SMPs on upstream edge of crest.
116 mm (0.07 to 4.07 116 mm to 11/56 (4 165 mm to 11/56 (4 yrs) @ Rapid deformation on first filling with significant reduction in rate thereafter. Crest
Lower Bear No. 2 - 0.271 73 0.171 0.128
years, to 11/56) yrs), ff 88 mm 15 m bc, ff 137 mm SMPs on upstream edge of crest.
Appendix E Page E8

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 5 of 8)
GENERAL DETAILS CONSTRUCTION DESIGN
Construction Timing Dimensions Facing Details (Support Zone for Face Slab)
Name Location /
Geology
Foundation
Main Rockfill (Comments) Thickness References
Owner Year Time Height, Length, Slab Face Upstream Downstream (Soil/Rock) Material Type /
L/H (normal to face Comments
Completed (years) H (m) L (m) Area (m2) Slope Slope Compaction
slab), m
high densities Knoop & Lack (1985), Knoop (1982b)
Greywacke, slate & finer graded greywacke
Mackintosh Hydro Tasmania 1981 2.75 75 465 6.2 27500 1.3 to 1 1.3 to 1 Rock Zone 3B not used. 3.7 rockfill, well compacted
achieved (higher than HEC (1991b), Bowling (1978)
phyllite main rockfill) Giudici et al (2000)
Mangrove Creek (Zone
3A) Interbedded Zone 3A - upstream facing (20 to 30%). Random fill Mackenzie & McDonald (1985)
Sydney Water, finer graded siltstone
sandstones, siltstones 1981 2 80 380 4.8 - 1.5 to 1 1.6 to 1 Weathered Rock (Zone 3B) makes up most of rockfill. Drainage layer 1.5 to 2 rockfill, well compacted
- PWSD NSW (1998)
Mangrove Creek (Zone NSW below random fill. Heinrichs (1996)
and claystones
3B)
finer crushed rockfill, Knoop & Lack (1985), Knoop (1982a)
Zone 3A - 70% of main rockfill, Zone 3B in outer Vey high strength
Murchison Hydro Tasmania Rhoylite 1982 2.25 94 200 2.1 17000 1.3 to 1 1.3 to 1 Rock 3.7 watered and well Gerke et al (1995), Robinson (1979)
downstream shoulder. rhyolite
compacted Giudici et al (2000)
Knoop & Lack (1985), Li (1991)
Zone 3A - 75% of main rockfill, Zone 3B in outer
laminated quartzite & Gravels in river crushed dolerite, finer Li et al (1991), Davies et al (1995)
Reece Hydro Tasmania 1986 3 122 374 3.1 37800 1.3 to 1 1.4 to 1 downstream shoulder. River gravels left in place (except 3.7 -
amphibolite section near upstream face).
grading, well compacted Partyka & Bowling (1984)
Giudici et al (2000)
some rock pieces
Pacific Gas & Derrick placed rockfill with Steele & Dreyer (1939)
Very coarse rockfill, "lot of large rocks with few fines". were observed to
Salt Springs Electricity Co. granite 1931 3 100 396 4.0 35300 1.3 to 1 1.4 to 1 Rock Likely low Cu .
4.6 flat face in the plane of the Steele & Cooke (1958), Regan (1997)
crush during
San Francisco face slab. Cooke & Strassburger (1988)
construction.

Salvajina (Zone 3A) Rock and soil, Alluvials in gully were in dense state (moduli approx 320
Tertiary sedimentary - gravels, thin layers, well Amaya & Marulanda (2000)
Columbia 1984 1.5 to 2 148 362 2.4 57500 1.5 to 1 1.4 to 1 gravels in river MPa). 2.8
siltstone and sandstone Zone 3A - upstream 2/3, Zone 3B - downstream 1/3.
compacted. Hacelas et al (1985), Sierra et al (1985)
Salvajina (Zone 3B) section

No Zone 3B. Bituminous concrete facing. Zone 2B (well-


interbedded argillite & sound crushed rockfill, Gerke et al (1993), Cole (1974)
Scotts Peak Hydro Tasmania 1972 1 43 1067 24.8 - 1.7 to 1 1.3 to 1 Rock compacted gravels) placed at upstream toe to 13 m 0.3 dolomite, well compacted.
-
Baker (1972), Cole & Fone (1979)
sandstone above foundation.

Segredo (Zone 3A) Zone 3A - upstream 1/3. Zone 3B (1C and 1D)
Basalt & Basaltic fine graded sound crushed high densities, higher Pinto et al (1985b), Pinto et al (1993)
COPEL, Brazil Breccia 1992 2.5 145 720 5.0 86000 1.3 to 1 1.4 to 1 Rock downstream 2/3. Dumped basalt rockfill in river section 3.1 basalt, well compacted than main rockfill Sobrinho et al (2000), Blinder et al (1992)
in downstream 2/3.
Segredo (Zone 3B)

Rock, gravels in No Zone 3B used. Grading of 3A indicates significant Forza & Hancock (1993)
Serpentine Hydro Tasmania quartzite and schists 1971 2 38 134 3.5 8000 1.5 to 1 1.5 to 1 0.3 150 mm minus rockfill -
river section breakdown on compaction. Giudici et al (2000)

Similar rockfill for entire embankment. Localised finer graded rockfill, well heavily compacted in
Shiroro Nigeria Granite 1983 - 125 560 4.5 - 1.3 to 1 1.4 to 1 Rock 6.1 Bodtman & Wyatt (1985)
enlargement of Zone 2A below EL 290 (bottom 30 m). compacted thin layers.

Tianshengqiao - 1
Sedimentary - crushed limestone (SW to Wu et al (2000a, 2000b)
(Zone 3A) Zone 3A - upstream half. Zone 3B (mudstone) high densities,
China limestone, sandstone 1999 3.3 178 1168 6.6 173000 1.4 to 1 1.3 to 1 Rock 4.6 FR), relatively fine, well Yang (1993)
surrounded by coarse limestone rockfill (Zone 3D). wetted.
Tianshengqiao - 1 and mudstone compacted. Jiyuan et al (2000)
(Zone 3B)
Knoop & Lack (1985)
finer crushed greywacke high densities
Tullabardine Hydro Tasmania Greywacke & slate 1979 0.33 25 214 8.6 5500 1.3 to 1 1.3 to 1 Rock No Zone 3B used. 3.7 Forza & Hancock (1995b)
rockfill, well compacted achieved
Bowling (1979)
1.5 to 2 fine crushed rock, thin Tuff and HEC (1987), Morse (1995)
White Spur Hydro Tasmania Volcanics - Tuff 1989 43 146 3.4 4300 1.3 to 1 1.3 to 1 Rock No Zone 3B used. 3.7
(?) layers, well compacted conglomerate. Morse & Ward (1989)
finer crushed rockfill, Regan (1980)
Melbourne siltstone with Zone 3A - upstream 1/3. Rockfill drainage layer below High densities in
Winneke 1978 1.4 85 1050 12.4 - 1.5 to 1 2.5 to 1 Rock 2.8 watered and well Casinader & Watt (1985)
Water interbedded sandstone random fill (Zone 3B) in downstream 2/3. Zone 2B
compacted MMBW (1975, 1981, 1995)
Pacific Gas &
derrick placed rockfill, 0.7 to Irregular surface Cooke (1958)
Wishon Electricity Co. Glaciated granite 1958 1.75 90 1015 11.3 60400 1.15 to 1 1.4 to 1 Rock Thinner rockfill lifts placed adjacent to upstream face. 2.4 to 3.5
2 m size, voids chinked finish Regan (1997)
San Francisco
Fresh crushed limestone, Peng (2000), Wang et al (1988)
Rock, gravels in high densities
Xibeikou China Sedimentary - limestone 1989 3 95 222 2.3 23000 1.4 to 1 1.4 to 1 Zone 3A - upstream half, Zone 3B - downstream half. 4.1 finer graded than Zone 1, Huang et al (1993)
lower river bed. well compacted
achieved.
Wang et al (1993)

Xingo (Zone 3A) Sobrinho et al (2000), Souza et al (1999)


fine crushed rockfill placed
Metamorphic - granitic Rock, gravels in Zone 3A - upstream 1/2; Zone 3B - downstream 1/2 high densities and Eigenheer & Souza (1999)
CHESF, Brazil 1993 6 140 850 6.1 135000 1.4 to 1 1.3 to 1 4.5 to 5.8 in thin layers, well
gneiss, Precambrian lower river bed. Dumped rockfill at downstream toe. low void ratios. Saboya et al (2000)
Xingo (Zone 3B) compacted.
Vasconcelas & Eigenheer (1985)
Appendix E Page E9

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 6 of 8)
Material Parameters / Properties of Rockfill
Rockfill Zone Intact Rock Strength Particle Size Distribution Parameters (from average grading curve) Density Placement Methods
Name Rockfill Particle
Wet/Dry % finer Dry Void Layer
n Authors Source Shape Strength Absorption D10 D30 D 50 D 60 D 80 D90 Dmax % finer Clean / Porosity, Level of
Class UCS (MPa) Strength Cu Cc 75 Density Ratio, Thickness Water Added Plant
Classn Classification (%) (mm) (mm) (mm) (mm) (mm) (mm) (mm) 19 mm Dirty n (%) Compaction
(%) micron (t/m 3) e (m)
Greywacke, some angular,
Mackintosh 3A 3A slate elongated
Medium to High 45 - - 52 3.3 1 13 34 52 130 230 1000 3.5 38 dirty 2.20 0.24 19.4 1.0 10% by volume 8p 10t SDVR Good

Mangrove Creek (Zone fresh siltstone & angular to


3A 1B sandstone subangular ?
High 45 to 64 40 - 60% 3.0 - 6.7% 310 9.9 0.45 25 88 140 225 280 400 4 27 dirty ? 2.24 0.18 15.5 0.45 to 0.6 5 to 7.5% by volume 4p 10t SDVR Good
3A)
weathered to fresh
Mangrove Creek (Zone siltstone &
angular to dirty (soft 2.06 Placed slightly dry of
3B 4 subangular ?
High 26 to 64 40 - 60% 3.0 - 6.7% 330 7.5 0.3 15 60 100 200 275 450 3 32 0.26 21 0.45 4p 10t SDVR Good
3B) sandstone
rock) t/m3 optimum

Rhyolite (SW to
Murchison 3A 3A FR)
angular Very High 148 - - 19 2.1 5.3 33 75 100 180 290 600 1.5 22 clean (?) 2.27 0.234 19.6 1.0 20% by volume 8p 10t SDVR Good

171 (80 to
Reece 3A 3A Dolerite angular Very High - - 10 1.6 16 63 120 160 330 475 1000 1 11 clean 2.287 0.29 22.7 1.0 5 to 10% by volume 4p 10t SDVR Good
370)

low 1000 2000 to sluiced, but poor dumped and Poor to Very
Salt Springs 3A 3A granite, Mesozoic angular, blocky Very High 100 to 130 - - - - - - - - low low clean (1.88) 0.41 29 5 to 52
(<10) (approx) 3000 quality poorly sluiced Poor

Salvajina (Zone 3A) 3A 2 natural gravels rounded (Very High) - - - 9.2 1.2 5.2 17 34 48 80 110 400 0 to 13 32 dirty 2.24 0.25 20 0.6 Yes, water added 4p 10t SDVR Good
weak sandstone
Salvajina (Zone 3B) 3B 4 angular ? Medium (?) - - - 45 3.4 1.2 15 41 58 115 190 600 0 to 16 32 dirty 2.26 0.21 17.4 0.9 Yes, water added 6p 10t SDVR Good
and siltstone

argillite - quarried angular ?


22 dry, 13
Scotts Peak 3A 3A Medium 60% - 380 6.6 0.21 10.5 44 80 205 370 914 7 38 dirty 2.095 0.266 21 0.915 no water added 4-6p 10t SDVR Good
wet

basalt (<5% Basalt = 235


Segredo (Zone 3A) 3A 1B basaltic breccia)
angular Very High - - 7.4 1.4 - - 177 - - - - - - clean 2.13 0.37 27 0.8 25% by volume 6p 9t SDVR Good
Breccia = 37

basalt (<5% High to Very Basalt = 235 2.01 1D - 0.8


Segredo (Zone 3B) 3B 1C & 1D angular - - 10.2 1.4 - - - - - - - - - clean 0.43 31 no water added 4p 9t SDVR Reasonable
basaltic breccia) High Breccia = 37 t/m3 1C - 1.6

Ripped quartz angular to (Medium to Not sure, suspect


Serpentine 3A 3A schist subangular
- - - 210 0.11 0.025 0.12 1.5 5.3 43 76 152 24 69 dirty 2.10 0.262 21 0.6 to 0.9 4p 9t SDVR Good
High) likely

Shiroro 3A 2 granite angular (Very High) - - - 32 2.1 4 33 95 128 260 380 500 - 22 clean 2.226 0.20 17 1.0 15% by volume 6p 15t SDVR Good

Limestone, SW to
Tianshengqiao - 1 15 to 90 to
3A 3B FR - spillway angular Very High 70 to 90 (wet) - - - - - - - - 800 - clean 2.19 0.23 19 0.8 20% by volume 6p 16t SDVR Good
(Zone 3A) excav.
20 120
Tianshengqiao - 1
3B 3C Mudstone angular Medium 16 to 20 - - 40 - - - - - - - 600 - 20 to 35 dirty 2.23 0.21 17.5 0.8 20% by volume 6p 16t SDVR Good
(Zone 3B)
Greywacke, some angular,
Tullabardine 3A 3A High 45 - - 28 1.7 2.8 19 50 78 155 240 400 2.5 30.5 dirty 2.22 0.23 19 0.9 to 1.0 > 10% by volume 4p 10t SDVR Good
slate elongated

Quarried Tuff - SW 37 to 0.18 to SDVR (4p, 10t


White Spur 3A 3A angular (Very High) - - - - - - - - - - 1000 - - - 2.30 15 to 20 1.0 > 10% by volume Good
to FR 200 0.25 ?)
SW to FR suspect 4 - 6p 10t
Winneke 3A 3 angular High 66 - 1.5 - 2.3% 33 1.7 2.8 21 58 91 200 310 800 4 28 2.07 0.302 23 0.9 15% by volume Good
Siltstone, Quarried dirty SDVR
Sluiced with high
550 to 1500 to 8 to 52 dumped and
Wishon 3A 3A granite, blasted angular Very High - - - (<7) - - - - - - - - clean ? 1.80 0.47 32 pressure jets, 300% Poor
750 2000 (variable) sluiced
by volume

Xibeikou 3A 1 Limestone - Fresh angular Very High 240 28 - - - - - 90 - - - 600 - - clean ? 2.18 0.284 22.1 0.8 25 to 50% by volume 8p 12t SDVR Good

granite gneiss angular


High to Very clean to
Xingo (Zone 3A) 3A 3 - - - 18 1.6 10 54 150 180 400 500 650 <3 4 to 33 2.15 0.28 21.8 1.0 15% by volume 4p 10t SDVR Good
High (?) dirty
Sound and
Medium to Very 0.3 to
Xingo (Zone 3B) 3B 4 weathered granite angular - - - 80 - - 190 - - - 750 2 to 7 15 to 60 dirty 2.1 0.31 23.6 2 no water added 6p 10t SDVR Reasonable
gneiss
High (?) 1
Appendix E Page E10

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 7 of 8)
SECANT MODULI VALUES HYDROGEOLOGY LEAKAGE
Modulus During, Er c (MPa) Modulus on First Filling, Erf Time for First Filling Rainfall,
Name Maximum
(from end of main average Long-term
Average at EOC Reservoir Level / Comments Rate Comment
Average over rockfill annual (l/sec)
(adjusted for Er f (MPa) Comments (l/sec)
construction period construction) (mm)
valley shape)
45 Face slab profile shows peak deformation is 24% above 8 to 10 (10 years
Mackintosh 39 63 toe, this is quite low.
1.92 to 2.93 years. Fluctuating (FSL to -10 m) 2135 21 Max. at end of first filling. Gradual reduction with time.
(35 to 60) post ff)
Mangrove Creek (Zone
55 to 60 57.5
3A) 0.6 to > 15 years (had not 2.5 (15 years
- Not reached FSL.
reached FSL by 1996).
First Filling - Oct 81 to ….. - 5.6 Max. leakage when storage level was at its highest (1991)
Mangrove Creek (Zone post EOC)
46 (36 to 56) 46
3B)

190 560 (485 to Large fluctuations on daily/weekly basis, up 2 (1995/96, 13


Murchison 154 - 1.04 to 1.09 yrs.
to 20 to 22 m.
- 3.5 3.5 l/sec on first filling.
(170 to 205) 640) years after ff)

190 (175 to 1.32 to 1.46 yrs (Apr to Fluctuating on daily/weekly basis, up to 8 0.5 to 1.5 (8 to 9
Reece 86 (57 to 115) 72 -
June 1986). m.
- 8 to 12 Initial leakage virtually all from right abutment.
205) years post ff)

Leakage mostly through the upper part of the face slab (upper 35
Calculated from face slab displacements measured after
Salt Springs - - 20 2 years
0 to 1.48 years (1931/32). Fluctuating 1145 565 - m) through cracks in face slab, honeycombe pockets in concrete
and open joints.

205 (175 to 260)


Salvajina (Zone 3A) 130
(likely affected by arching) By He (2000). GJH estimate is 615 MPa when reservoir Approx. 14 l/sec through abutments and 60 l/sec through face
500 at 14 m below FSL.
0.33 to >0.75 years (1985). no details after first filling - 74 - slab.
Salvajina (Zone 3B) 62 (likely affected by arching) 53
Estimated from HSGs near to upstream face.
59 (Zone 3A) Significantly higher modulus for gravels in lower section
0.42 to 2.58 years (June 72
Steady
2 to 4 (1994, 20 At close to FSL on first filling leakage increased from 5 to 100
Scotts Peak 20.5 (18.5 to 23.5) 20.5 to Aug 74).
2130 100 l/sec due to cracking in face slab. Gravel blanket placed.
420 (Zone 2B) compared with argillite rockfill.
years post ff)

55 (excludes points thought to


Segredo (Zone 3A) 55
be affected by arching) 45 (1.5 to 2 Max leakage on first filling when at FSL. Leakage reduced by
175 - Approx. 0.6 years. - - 390 years post ff) dumping fill on lower portion of face slab.
Segredo (Zone 3B) 28 (25 to 33) 28

Assumption that the river gravels have a significantly


0.24 to 2.94 years (Dec 71 No estimates possible due to inundation of the downstream toe of
Serpentine 92 (46 to 142) 92 97 (94 to 100) greater modulus (than rockfill) and have negligible
to Aug 74).
Steady - - - the embankment
contribution to deformation on first filling.

No deformation quoted for upstream face slab on first 1.44 to >1.8 years (from 100 (0.5 years When close to FSL. Cracking in face slab near junction with
Shiroro 66 (61 to 71) 58 - By Oct 1984 not a FSL (10 m short) - 1800
filling. 5/84). after max.) plinth.

Tianshengqiao - 1
49 (40 to 57) 49
(Zone 3A) Not yet filled to FSL. Stored water during construction First filling started 1.3 years before end of Max at end Dec 1999 when reservoir at highest level. March
- -1.5 to >0.8 years. - 53 -
to within 40 m of FSL. construction. 2000 ~ 30 l/sec. Repairs to cracks in face slab.
Tianshengqiao - 1
37 (32 to 42) 37
(Zone 3B)
2.05 to 2.35 years (5/81 to 0.5 to 1 (10 -12
Tullabardine 74 74 170 - Fluctuating (FSL to -10 m) - 1.5 to 2 Current base flow less than 1 l/sec.
8/81). yrs post ff)
180 On virtually date of reaching FSL. 1 year later E = 200 to 0.18 to 0.24 years (June to 2 (6 years post Limited leakage. &l/sec shortly after first filling. Gradual
White Spur 139 340 Fluctuating 3150 7
(160 to 200) 240 MPa. July 1989). ff) reduction to approx. 2 l/sec.

1.62 to 5.04 years (6/80 to Fluctuating - seasonal (no catchment, off 13 (10 to 12 Max. leakage when reservoir at FSL on first filling. Gradual
Winneke 55 (50 to 59) 55 104 Determined from HSGs close to upstream face. 700 to 1200 58
11/83). stream storage) years post ff) decrease in leakage rate since.

Maximum within 2 months of first filling. Due to cracking in face


0.4 to 0.45 years (May
Wishon - - - - - - 3120 850 slab, mainly at the perimeter joint. Series of repais to face slab
1958).
over the years.
Impounded water during construction.
Query on deflected shape of face slab after September -1.5 to 6 years (June
Xibeikou 80 (60 to 100) 80 260 Overtopped in 1987. Other problems 1150 - - -
1993 1995).
during construction.

Xingo (Zone 3A) 34 (30 to 39) 34 Steady. Springs observed on downstream


Calculated for measurements taken on the left abutment.
1.02 to 1.46 (mid to late
slope of left abutment (associated with
140 (4.5 years Leakage increased post ff from 100 to 200 l/sec. Due to cracks in
76 (73 to 80) 1994) years.
- 200 slab. Dumping soils on face reduced leakage to 140 l/sec.
seepage through the face slab)
post ff)
Xingo (Zone 3B) 13 (12 to 14) 13
Appendix E Page E11

Table E1.1: Case studies information on concrete face rockfill dams (Sheet 8 of 8)
MONITORING
Settlement Post Construction Crest Deformation Long Term Crest
Deformation Normal
Name During
Settlement Settlement Lateral to Face Slab
Settlement Rate
Comments
Other Comments
Construction Settlement, total Settlement, (% per log time cycle)
on FF on FF Displacement (mm)
(mm) (mm) total (%) (t o at EOC)
(mm) (%) (mm)
333 mm (0 to 20.6 Decreasing rates of deformation with time, approximately 5 mm/yr settlement and 5
Mackintosh > 780 0.444 99 0.132 130 mm (75 ff) 228 mm (173 ff) 0.184 to 6 mm/yr displacement at 10 yrs.
years)
Mangrove Creek (Zone
3A) 0.251 (after 5 years, but still First monitoring of crest SMPs 0.67 years after end of rockfill construction.
287 mm (0.67 to 15 196 mm (from 12/81 Moduli during construction calculated from HSG deformation during construction and also
> 610 0.359 > 287 > 0.359 to 5/96)
- not reached full supply Settlement rates with time still high to 1990 (highest water level) then reduced from
taken from published values.
Mangrove Creek (Zone yrs, 12/81 to 5/96) 1990 to 1996.
level)
3B)

104 mm (0.08 to 17.6 77 mm to 4/93 (28 ff 67m


Murchison > 255 0.111 9 0.01 22 mm to 4/93 (8 ff) 0.056 Decreasing rates of movement with time, approximately 3mm/yr at 10 yrs
yrs, to 1999) bc)

221 mm (0.12 to 15 68 mm (5/86 to 264 mm to 8/94 (215 ff Decreasing rates of movement with time, approximately 4 mm/yr crest settlement
Reece > 820 0.181 85 0.07 11/93), all post ff
0.063 at 5 yrs
yrs, to end 1999) 66m bc (45%))

Rapid deformation on first filling with significant reduction in rate thereafter.


1276 mm to 1996 550 mm to 1958 (27 1755 mm to 1958 (67 m
Salt Springs - 1.276 380 0.38 yrs), ff 230 mm
0.29 to 0.77 Acceleration at about 20 years when the full supply level was raised by 3 m. Crest
(0.26 to 65 yrs) bc), ff 1317 mm SMPs on upstream edge of crest.

Salvajina (Zone 3A)


90 mm (0.33 to 0.75 55 mm to 14 m short of Constructed of dirty gravels with rockfill used in the downstream 1/3. High stiffness as
- 0.061 > 90 > 0.061 - - Limited data.
indicated by high moduli. Limited deformation data in published literature.
years) FSL
Salvajina (Zone 3B)

445 mm (0 to 18 190 mm to 12/89 178 mm to 12/89, 78 mm High rate of deformation post first filling due to leakage through dam and wetting up Cracking occurred as a result of tensile stresses in face slab due to very large differential
Scotts Peak - 1.036 203 0.472 (17 yrs), ff 96 mm
0.178 (after 5 years) of argillite rockfill. Long term rates approximately 2.5 mm/yr. stiffness in rockfill.
years, to 12/89) ff

Segredo (Zone 3A)


229 mm (to 0.4 yrs 340 mm (to 4 months Limited information on deformation. Estimation of crest settlement from HSG
- 0.158 Approx. 200 0.138 - -
post ff, 1992) post ff) device below crest.
Segredo (Zone 3B)

77 mm (0.24 to 25.5 38 mm (12/71 to


Serpentine - 0.203 35 0.092 1/92), ff 23 mm
- 0.109 -
yrs, to 1/97)

Modulus during construction taken from Bodtman and Wyatt (1985), however they may
166 mm (0 to 1.8 27 mm to 12/84 (1 90 mm (when 32 m Limited deformation information. Could find nothing on deformation behaviour after
Shiroro - 0.133 > 66 > 0.053 - not have allowed for reduction in the vertical stress due to embankment shape.
years) to 12/84 yr) below FSL) 1984.
Therefore, the values are likely to over-estimate the modulus.

Tianshengqiao - 1 During the early stages of construction the embankment (~40 m high) was overtopped 4
670 mm to Dec
(Zone 3A) 926 mm (0.05 to 0.8 times during floods. Cracks noted in slab cushion zone during construction. Face slab
- 0.52 > 926 > 0.52 1999 (0.75 years), ff - - First filling not completed.
years, to Dec 1999) not completed.
cracking on first filling. Separation between face slab and cushion zone was noted during
Tianshengqiao - 1 staged construction of the face slab, up to 100 mm gap.
(Zone 3B)
19 mm (0.2 to 12.8
Tullabardine - 0.076 2 0.01 - - 0.023 Negligible post construction settlement
years)
58 mm (0.04 to 5.9 38 mm (@ 6 years), 15 Rate of settlement reduced after completion of wave wall in January 1990. At 5yrs
White Spur > 65 0.135 7 0.016 - 0.08
yrs, 4/89 to 2/95) mm ff crest settlement rate approx. 1.25 mm/yr.
207 mm (0.17 to 16.2
M7 - 160 mm (3/80 to 2/94), Monitoring started shortly after end of rockfill construction (SMP 13 on upstream
Winneke - yrs, Jan 79 to Jan 0.244 105 0.124 - 0.107
145 mm ff (all settlements) face used until crest SMP installed). Deceleration of settlement after first filling.
95)
Monitoring started shortly after end of rockfill construction. Rapid movement on
Wishon - 954 mm (0 to 38 yrs) 1.136 189 0.233 - - 0.25 to 0.33 No data to calculate modulus during construction or on first filling
first filling with significant reduction in rate thereafter.

Limited data available. Monitoring shows discrepancies in face slab deformation, and
Xibeikou - - - - - - 75 mm (6.5 years), all ff - Query on results. Discrepancies between SMPs and HSGs in published literature.
between surface monitoring points and hydrostatic settlement gauges.

Xingo (Zone 3A) SMPs base survey approx. 1 year after the end of rockfill construction. Cracking of Wetting of rockfill in left abutment resulted in acceleration in the rate of deformation (in
320 mm (1.0 to 6.2 510 mm (@ 6 years), face slab on left abutment and subsequent leakage resulted in acceleration of this area). Springs observed high above river level on downstream slope of embankment
- 526 (1.0 to 6.2 yrs) 0.376 302 0.216 yrs), 210 mm ff
0.25
290 mm on ff. deformation of face slab and crest (i.e. wetting of rockfill). (on the left abutment) indicative of wetting of Zone 3B rockfill.
Xingo (Zone 3B)
APPENDIX F

Summary Tables and Plots for


Embankment Dam Analysis
Chapter F Page i

TABLE OF CONTENTS

1.0 SUMMARY TABLES FOR CASE STUDIES ............................F1

2.0 SUMMARY PLOTS OF POST CONSTRUCTION


DEFORMATION BEHAVIOUR ..............................................F61
2.1 Post Construction Deformation of the Downstream Slope ...........................F61
2.2 Post Construction Deformation of the Upper Upstream Slope to
Upstream Crest Region .................................................................................F76

LIST OF TABLES
Table F1.1: Central core earth and rockfill embankments case study information.........F5
Table F1.2: Zoned earth and rockfill embankments case study information................F35
Table F1.3: Zoned earthfill embankments case study information...............................F45
Table F1.4: Earthfill embankments (homogeneous earthfill, earthfill with filters
and earthfill with rock toe) case study information................................................F55
Table F1.5: Summary of embankment properties of puddle core earthfill
embankments..........................................................................................................F58
Appendix F Page F1

1.0 SUMMARY TABLES FOR CASE STUDIES


Table F1. to Table F1.5 present a summary of aspects of the design, construction,
embankments materials, reservoir operation, performance and references for most of the
earthfill, zoned earth and earth-rockfill and puddle core earthfill embankment case
studies within the database. The five tables are:
• Table F1. (of 30 pages) – central core earth and rockfill embankments;
• Table F1.2 (of 10 pages) – zoned earth and rockfill embankments, that are not of
central core earth and rockfill design;
• Table F1.3 (of 10 pages) – zoned earthfill embankments;
• Table F1.4 (of 3 pages) – earthfill embankments including homogeneous, earthfill
with filters and earthfill with rock toe embankment types; and
• Table F1.5 (of 3 pages) - puddle core earthfill dams.

For embankments with rolled earthfill zones (i.e. earthfill, zoned earthfill and zoned
earth and rockfill embankments), the embankment performance aspects include a
summary of the deformation behaviour of the core during construction, post
construction deformation of SMPs (surface monitoring points) on the crest and
shoulders (mainly from close to the maximum section), the positive pore water pressure
response in the main earthfill zone (both during and post construction) and comments
on observed cracking and overall observed performance.
For the puddle core earthfill embankments, the hydrogeology is an assessment of the
response of the various embankment zones (upstream shoulder, puddle core and
downstream shoulder) to changes in the reservoir level. Where possible the comments
have been based on actual piezometer and standpipe records, however, for a number of
embankments the response has been assumed based on assessment of the earthfill type
and the presence or otherwise of a select earthfill zone in the upstream shoulder. The
deformation monitoring records for the puddle core earthfill dams are fairly limited
given the age of most of the dams in the data set. For most dams the monitoring records
are from SMPs typically established in the 1970’s to 1980’s, many 10’s of years after
completion of construction. Some monitoring data and anecdotal information is
available on the early performance of several embankme nts for estimation of the
settlement from the end of construction.
Appendix F Page F2

Explanations and definitions of the terminology used in the tables (mainly for
Tables F1.1 to F1.4, but also applicable to Table F1.5) is:
• General terminology and acronyms used throughout:
− The earthfill materials have been classified in accordance with the Australian
soil classification system (Australian Standard AS 1726-1993 Geotechnical Site
Investigations).
− The strength and weathering terms used to describe the rock used as rockfill are
from the case study references. The acronyms FR = fresh, SW = slightly
weathered, MW = moderately weathered, HW = highly weathered.
− EOC = end of construction
− FF = first filling
− FSL = full supply level
− Fndn = foundation
− PWP = pore water pressure
− UCS = unconfined compressive strength
− SMDD = Standard (or Standard Proctor) Maximum Dry Density
− MMDD = Modified Maximum Dry Density
− OMC = Optimum moisture content for the laboratory compaction test method
used. In most cases it is Standard or Standard Proctor optimum, but in several
instances modified compaction or an adjusted compaction test type have been
used.
− Unk or nk = unknown. In most cases a “–” has been used, which indicates either
the data is not known or is not appropriate.
• Instrumentation terminology is:
− SMP = surface monitoring point
− IVM or ES = internal vertical settlement gauge
− IHM = internal horizontal monitoring gauge
− IN = inclinometer
− HSG = hydrostatic settlement gauge
• Terminology used for the owner / authority is:
− USBR = United States Department of the Interior, Bureau of Reclamation
− SMHEA = Snowy Mountains Hydro Electric Authority
− State Water DLWC = Department of Land and Water Conservation
Appendix F Page F3

− ACTEW = Australian Capitol Territory Electricity and Water Corporation


• For the construction / design section:
− The embankment zoning classification system used is defined in Section 1.3.3 of
Chapter 1.
− Length, L = embankment crest length.
− The embankment slopes are defined in terms of horizontal (H) to vertical (V).
• For the embankment materials section:
− ASCS = Australian soil classification system
− For placement methods, 6p 18t roller = 6 passes of an 18 tonne roller. SDVR =
smooth drum vibrating roller.
− DR = relative density.
− The “level of compaction” is a qualitative rating system of the degree of
compaction. Details for the rockfill classification are given in Section 1.3.3 of
Chapter 1.
− The “Clean / Dirty” classification for rockfills is also a qualitative rating system.
“Dirty” rockfills refer to those with a high fraction less than 19 to 25 mm.
Generally, 25% has been used to delineate between clean and dirty rockfills, but
if the references indicate the rockfill to be “dirty” then it is classified as such.
• For the hydrogeology section:
− “Fluctuating” is where the reservoir is subject to a seasonal (usually annual) or
regular (more than once per year) drawdown that is typically greater than about
0.1 to 0.2 times the height of the embankment at its maximum section.
− “Fluctuating slow” is where the reservoir is subject to fluctuations that occur
slowly over time (i.e. slow drawdown over several years).
− “Steady” is where the reservoir remains steady over time (i.e. where the
fluctuation is less than 0.1 to 0.15 times the height of the embankment at its
maximum section).
− ru = pore water pressure coefficient (refer Equation 7.4 in Section 7.4.1.4 of
Chapter 7)
• For monitoring during construction (Table F1. to Table F1.4):
− The core settlement refers to the total settlement of the core during the period of
construction and is given in units of millimetres and as a percentage of the
embankment height, H, at the point of measurement.
Appendix F Page F4

− If no information is given in the section “Other Types of Internal Monitoring


Equipment” is does not necessarily mean no other deformation monitoring
instrumentation was installed for that embankment.
• For monitoring post construction (Table F1. to Table F1.5):
− The zero time reading is at the end of construction. Time readings given in this
section are in terms of years after the end of construction (e.g. 1.3 years = 1.3
years after end of construction).
− Settlements and displacements are generally from SMPs on or close to the main
section. Values of settlement in % are settlement as a percentage of the height
from the SMP to foundation level.
− SLT = long-term crest settlement rate in units of settlement (as a percentage of
the height from the SMP to foundation level) per log cycle of time.

Comments specific to the puddle core earthfill embankment case studies (Table
F1.5) are:
• In the classification of the puddle core, the terms “low”, “medium” and “high” refer
to the soil plasticity.
• For the post construction deformation:
− The total crest settleme nts since end of construction are given in terms of
millimetres and as a percentage of the embankment height, H.
− The long-term settlement rates are, in most cases, specific to the crest settlement
unless otherwise stated. The quoted values are for periods of normal reservoir
operating conditions.
Appendix F Page F5

Table F1.1: Central core earth and rockfill embankments case study information
GENERAL DETAILS CONSTRUCTION / DESIGN
Embankment
Construction Timing Dimensions
Classification
Name Owner/ References
Location Geology Foundation Dam Core Upstream Downstream Comments on Design
Authority Year Time Core Height, Length, Ratio
Zoning Width / Slope Slope
Completed (years) n Type H (m) L (m) L/H
Class Slope (H to V) (H to V)

Water Welded tuffaceous


Thin central clay core supported by compacted rockfill
Resources rhyolite with granitic Karasawa et al (1994)
Agigawa Japan Rock 1990's - 5,1,1 c-tn CL 102 460 4.5 2.6H to 1V 2H to 1V shoulders. Single filter zones up and downstream of core
Development porphry intrusions, highly Yamazumi et al (1991)
and drainage layer on downstream foundation.
Public Corp. fractured.

Very narrow central core with near vertical downstream edge


River section of embankment on
Ajuare Sweden - Glacial activity. 1966, Oct 2 5,2,2 c-tn SM 46 525 11.4 1.8H to 1V 1.7H to 1V supported by poorly compacted rockfill shoulders. Nilsson & Norstedt (1991)
bedrock. Abutments on glacial till.
Foundation filters used due to high fines content of rockfill.

Central clay core aligned slightly to upstream supported by


Cetin et al (2000)
rockfill shoulders of weathered to fresh rock types. Deep
Cetin (2002)
Ataturk Turkey - Limestone Rock - karst limestone 1990, Aug 3.6 5,2,0 c-tm CH 184 1664 9.0 2.15H to 1V 2.2H to 1V grout curtain (>180 m) in karst limestone. Cofferdam
Oziz et al (1990)
incorporated into the upstream toe of the embankment. The
Lask & Reinhardt (1986)
dam axis is arched to upstream.

Thin central core with rockfill sourced from weathered to


Sedimentary -
Bath County Upper fresh rock types. A zone of fresh rockfill is used immediately
USA, Virginia - sandstones and Rock 1984, Dec 3 5,2,0 c-tn SC (?) 134 670 5.0 2.4H to 1V 2.5H to 1V Wong et al (1992)
Dam downstream of the filters and on the downstream foundation
siltstones
for leakage control purposes.

Naylor et al (1997)
Core and filters on MW schists, rockfill Narrow central core with weathered (Zone 3A) to fresh (Zone
Maranha Das Neves et al (1994)
Beliche Portugal - Schists on alluvial sands and gravels of 1985, Mar 3 5,2,2 c-tn GC 55 527 9.6 2H to 1V 1.95H to 1V 3B) poorly compacted rockfill shoulders. Cofferdam
Naylor et al (1986)
medium density. upstream of the main dam.
Pagano et al (1998)

Medium width central core with dry dumped rockfill


Sedimentary, mainly Variable. At maximum section core
Wimmera shoulders. Embankment design changes associated with Currey et al (1968)
Bellfield Australia, Victoria sandstones; granite and upstream shoulder on sandstone, 1966, Apr 2.25 5,2,0 c-tm CL/SC 40 823 20.6 1.6H to 1V 1.42H to 1V
Mallee Water changes in foundation conditions. Slopes a series of steep SMEC (1998a)
intrusions. downstream shoulder on alluvium.
benches at 1.2 - 1.33H to 1V (angle of repose).

Complex geology, mix of QWRC (1986a, 1986b)


Australia, QDNR (now sedimentary, Excavation to bedrock for core and Very narrow central core (0.1H to 1V up and downstream) Eadie (1988)
Bjelke Peterson 1988, Sept 1.7 5,2,1 c-tn CL 41.5 650 15.7 1.7H to 1V 1.7H to 1V
Queensland Sun Water) metamorphic and rockfill shoulders. with compacted rockfill shoulders. QDNR (1997)
volcanic Hadgraft (1984)

SMHEA (1964)
Slightly metamorphosed Svenson (1964)
Australia, New State Water sedimentary - Excavation to bedrock for core and Central core (slopes of 0.5H to 1V up and 0.4H to 1V Olsauskas et al (1966, 1967a,
Blowering 1968, Apr 2 5,2,0 c-tm SC - CL 112 808 7.2 1.9H to 1V 1.9H to 1V
South Wales DLWC interbedded phyllite, rockfill shoulders. downstream) supported by compacted rockfill shoulders. 1967b, 1968)
siltstone & quartzite Hunter & Bacon (1970)
Bacon (1969, 1999)

Core and most of downstream shoulder Central clay core supported by compacted rockfill shoulders.
DWR NSW (1989)
Australia, New State Water Sedimentary - Jasper & on bedrock (MW to FR). Upstream Cofferdam at upstream toe. Large bench on upstream slope
Chaffey 1979, Mar 1 5,2,0 c-tm CH/GC 54 530 9.8 1.75H to 1V 1.75H to 1V WRC NSW (1979)
South Wales DLWC Siltstone shoulder on dense alluvial gravels (~ 12 at 24 m depth below crest level. Disposal area at
Newland & Davidson (1979)
thick). downstream toe.

Broad central core (0.7-0.75H to 1V) with thin filters zones


Hetch Hetchy Granite, glacially scoured Rock. Strip to weathered rock, cutoff to Lloyd et al (1958)
Cherry Valley USA, San Francisco 1955, Oct 2 5,2,0 c-tk SM/ML 100 793 7.9 1.85-2H to 1V 1.85-2H to 1V and dumped and sluiced rockfill shoulders. Outer slopes in a
Water Supply valley fresh granite. Cooke & Strassburger (1988)
series of benches sloped at 1.33H to 1V.

Core on bedrock. Shoulders in river


Narrow central core with wide filter zones supported by well
Limestone, some clay section on up to 60 m depth of granular Alberro & Moreno (1982)
Chicoasen Mexico - 1980, May - 5,2,0 c-tn GC 261 463 1.8 2.35H to 1V 2H to 1V compacted rockfill shoulders. Constructed in 125 m wide
shales alluvial deposits containing large Moreno & Alberro (1982)
gully with near vertical abutment slopes.
boulders.
Appendix F Page F6

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 2 of 30)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Relative
Layer
ASCS Ratio or Dry Moisture ASCS Density or Level of
Zone Source n Placement Methods Comment Zone Source n Thickness Plant Comment
Class Density Content Class Dry Density Compaction
3 3 (mm)
(t/m ) (t/m )
Limited information, suspect well compacted.
(bulk density = Well compacted (?) - heavy Likely well-compacted, typical practice in likely from crushed 3
Agigawa 1 - CL (?) 3 - 2A (filter) - 2.12 t/m - - - Very high moduli (Karasawa et al 1994). Single
2.10 t/m ) rollers (?) Japan is for heavy rolling of core. rock
filter up and downstream of the core.

Reasonably to well 4.5t SDVR, no indication of


(bulk density = Silty morraine core placed well wet of 2A & 2B (bulk density = reasonable to Dual filters used up and downstream, combined
Ajuare 1 glacial till SM 3 compacted (?) - heavy crawler 4 to 6% wet of OMC river alluvium - 3 500 number of passes of
2.24 t/m ) OMC using Swedish method. (filters) 2.14 t/m ) good width of 8 to 13 m.
type tractors rollers.

Reasonably to well compacted sandy


reasonably to well compacted - clays of medium to high plasticity. Cetin compacted by vibratory
CH - sandy 94.4% SMDD average 1.5% dry of 2A & 2B SP/SW (2A) 88 to 91% SMDD reasonable to Low fines content, less than 1.5%. Dual filters
Ataturk 1 - 3 300 mm layers, 6p sheepsfoot et al (2000) query layer thickness used, river alluvium 3 - rollers, no indication if
clays 1.52 t/m OMC (filters) GP/GW (2B) 1.84 - 1.97 t/m good (?) used up and downstream of the core.
roller they indicate it may have been up to 1m water added.
at times.

limited information. Dual filters downstream of


Core of stiff to very stiff strength
Bath County Upper soil and weathered 2A & 2B the core. Upstream of the core, dual filters in the
1 SC/GC (?) - no information - consistency, suggesting it is possibly well crushed limestone - - - - -
Dam rock (filters) upper section and a single filter in the mid to
compacted and placed at close to OMC.
lower section.

GC - clayey Suspect reasonably to well compacted limited information. Suspect at least reasonably
2A & 2B
Beliche 1 - sandy gravels, - no information close to OMC (Spec) given period of construction. Clayey gravels (2B) GW/GP (2B) - - - - compacted. Dual filter used up and downstream
(filters)
20 to 36% fines sandy Gravels with low plasticity fines. of the core.

CL/SC - sandy Reasonably to well Layer thickness on the large size for roller quartzose Compacted using SDVR, Limited information. Narrow width dual filters up
Spec 1.5% dry to 2A & 2B GM (2A)
Bellfield 1 alluvial (?) clays to clayey - compacted - 380 mm layers, size (in todays standards). Sandy clays sandstone (some - - no indication if water - and downstream of the core. Zone 2A contained
1.5% wet of OMC (filters) GP (3A)
sands 12p 4t sheepsfoot roller to clayey sands. sand and gravel) added. 5 to 15% fines.

Sandy clays to clayey sands of medium Spec >95% Fines content less than 15% (2A), less than 2%
CL, SC (some 100% SMDD Well compacted - 150 mm (Spec. 1% dry to 1% 2A & 2B alluvial sands & SM/SP (2A) 10t SDVR, moist at
Bjelki-Peterson 1 colluvial, alluvial ? 3 plasticity, some high plasticity sandy SMDD (2A) and 350 Good (2B). Dual filters downstream of core and single
CH) 1.72 - 1.88 t/m layers, pad foot rollers wet of OMC) (filters) crushed rock GP/GW (2B) compaction
clays. >70% DR (2B) (Zone 2A) filter upstream.

Clayey sands and sandy clays of medium


SC/CL - clayey
plasticity. Changes in moisture
sands to sandy Well compacted filter zones. Zones 2A and 2B
Colluvial, alluvial, Well compacted - 150 mm average 0.2% wet of specification as construction proceeded. 2A & 2B 2A - SW to GW 3
Blowering 1 clays, avg 46% 101.5% SMDD Alluvial 2.14 t/m 450 4p 7.5t SDVR Good used up and downstream of the central core.
residual layers, 8p sheepsfoot roller OMC Bottom 34 to 37m 0.3% dry of OMC; mid (filters) 2B - GW/GP
fines (30 to Combined width of approx. 6 m.
region 0.3% wet of OMC; upper 40m
80%)
0.1% wet of OMC.

CH/GC - gravelly
sandy clays to Gravelly sandy clays of medium to high Well compacted filters. Dual filters downstream
102.4% SMDD Well compacted - 150 mm 2, 2A & 2B D R = 88%
Chaffey 1 colluvial clayey gravels, 3 1.5% dry of OMC plasticity. 0.6 m CH contact layer. Core river gravels GP to GW 3 400 minimum 1p 8t SDVR Good of the core (5m width), single filter upstream (2.5
1.86 t/m layers, 3 to 4p, tamping rollers (filters) 2.10 t/m
average of 49% centreline aligned slightly to upstream. m width).
fines

SM/ML - silty Well compacted silty sand to sandy silt


Well compacted - 225 mm
decomposed sands to sandy 91% MMDD core. Fines of low plasticity (PI = 1 to 2A & 2B alluvial sands and GP/GW (sand 3 300 Thin (6 m wide) dual filters used up and
Cherry Valley 1 3 (loose) layers, 12p 31t - 1.76 t/m 1p crawler tractor. Reasonable
granite silts, 10 to 82% 1.70 t/m 5%). Likely placed wet due to wet field (filters) gravels to cobble size) (compacted) downstream of the core.
sheepsfoot roller
fines condition (5 to 15% wet of OMC).

Clayey gravels with medium plasticity


GC - clayey Well compacted - 250 mm fines. Higher clay content material used in GP/GW - sandy Dual filters up and downstream of the core.
residual soils 2A & 2B
Chicoasen 1 gravels, 12 to - layers, 6p 7t padfoot vibratory close to OMC lower portion. Placed 2 - 3% wet of OMC alluvial (?) gravel (2A) to - 400 2p 10t SDVR Good Filters zone moderately wide in the lower half of
(weathered lutite) (filters)
30% fines roller. in the zone against the steep abutment gravel (2B) the embankment.
slopes.
Appendix F Page F7

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 3 of 30)
MATERIALS
Rockfill Zone Rockfill Zone (in some cases earthfill stabilising berm)
Name
Dry Layer Level of Dry Layer Level of
Grading Clean / Grading Clean /
Zone Source Density Thickness Placement Method Comp- Comment Zone Source Density Thickness Placement Method Comp- Comments
Description Dirty 3 Description Dirty 3
(t/m ) (m) action (t/m ) (m) action

Suspect well compacted given typical


3
Agigawa 3A & 3B - - - 2.00 t/m - SDVR - procedures used in Japan at this time. - - - - - - - - -
Moduli = 130 MPa (Karasawa et al 1994).

Poorly compacted rockfill, and suspect


schists and (bulk density = Spread by D8 dozer then
Ajuare 3 high fines content dirty 3 - Poor relatively poor quality rock given propensity - - - - - - - - -
gneisses 2.04 t/m ) sluiced after placing.
to breakdown.

weathered basalt Both rock types of medium to high UCS. Suspect sound basalt of very high UCS.
compacted by vibratory compacted by vibratory
(upstream), high fines content 2.03 - 2.06 Weathered basalt used as inner upstream 3 Limited information on placement methods.
Ataturk 3A dirty 3 0.6 to 1.5 (?) rollers, water added if Reasonable (?) 3B basalt, sound - clean ? 2.26 t/m 0.6 to 1.5 (?) rollers, water added if reasonable (?)
limestone for rockfill t/m rockfill zone and pliacated limestone used Zone 3B predoninantly used in the outer up
moisture content < 2%. moisture content < 2%.
(downstream) as inner downstream rockfill. and downstream shoulders.

siltstone and max 760 mm, Limited information, suspect well-compacted.


Siltstone and Limited information, suspect well-
Bath County Upper max 760 mm sandstone, SW to < 10% fines (3A) 3A - dirty Zones 3A used as filter zone in the
3B sandstone - dirty - - - - compacted. Zone 3B is the predominant 3A, 3C - - - -
Dam 10 to 25% fines FR (3C sandstone <10% passing 5 3C - clean ? downstream shoulder and Zone 3C used in
weathered to FR. rockfill type used.
only) mm (3C) the upper upstream shoulder.

Lightly compacted. Zone 3A used as inner


rockfill zone up and downstream of the max 600 mm Lightly compacted. Zone 3B used as outer
schists and Light compaction, no Poor to Greywacke, high Lightly compacted, no Poor to
Beliche 3A significant fines dirty - 1.0 core. Poor quality weathered rock used 3B <20% < 19 mm clean ? - 1.0 rockfill zone up and downstream. Greywacke
greywacke, HW indication if water added. reasonable (?) strength indication if water added. reasonable
(Wet UCS = 9 to 12 Mpa, Moduli approx. 5 no fines of high strength (UCS = 30 to 45 MPa).
to 27 Mpa).

quartzose Dumped in 1.2 to 9.1 lifts,


Lift thickess typically 1.2 to 2.4 m.
sandstones no water added. 5 m alluvium and Limited information. At main section, Zone 4
Bellfield 3A - - - 1.2 to 9.1 poor Susceptible to significant collapse 4 (random) - - - - - -
(some siltstone width next to filters - 1.2m decomposed rock used as berm on the downstream shoulder.
compression given dry placement.
and mudstone) layers 6p 4.5t SDVR.

Phyllite (MW or max 300 mm 3 4p 10t SDVR, 20% by Well-compacted rockfill. Rock of medium to
Bjelki-Peterson 3A dirty Spec > 2 t/m 0.7 Good - - - - - - - - -
better) 4 to 12% fines volume water added high UCS.

Rock of high to very high strength when


Quartzite of high to very high strength when
fresh, with large reduction in UCS on wetting
Sluiced (0.9 to 1 ratio by dry, large reduction (approx 30%) on Sluiced (0.9 to 1 ratio by
900 (35 to 60%), particularly in the phyllite.
volume) when tipped, saturation. Zone 3A used as the inner volume) on tipping, then
max < 900 mm 3 (450 for 6m Quartzite & Max < 1800mm 3 Possible usage of lesser quality rock in lower
Blowering 3A Quartzite clean to dirty 2.06 t/m then spread by dozer and Good rockfill zone up and downstream of the 3B dirty 2.07 t/m 1800 spread by dozer and reasonable
18 to 28% < 25mm zone adjacent Phyllite 17 to 50% < 25mm elevations. Significant breakdown on surface
compacted by 4p 7.5t central core. Also used a more permeable compacted by 4p 7.5t
to filters) of layer and large density variation within
SDVR layer below Zone 3B in the downstream SDVR
layer. Zone 3B used in outer up and
shoulder.
downstream shoulders.

Spread by dozer, 2 to 4p
Jasper (SW to 26 to 57% < 19mm 3 reasonable to
Chaffey 3 dirty 2.02 t/m 1.2 8t SDVR, no indication if Some siltstone in rockfill. - - - - - - - - -
FR) %fines = 1 to 4%. good
water added.

Granite and
Rockfill adjacent to the filters was of lesser
Cherry Valley 3 granodiorite, - clean ? - 9.1 dumped and sluiced poor - - - - - - - - -
quality and was not sluiced.
fresh

max 500 to 600mm Zone 3B formed outer 10 to 25% of the


4p 12t SDVR, water Heavily compacted rockfill. Zone 3A
Chicoasen 3A limestone 5 to 50% < 19 mm dirty - 0.6 Good 3B limestone - - - - Dumped Poor rockfill zone of both the upstream and
added formed most of rockfill in shoulders.
0 to 10% fines downstream shoulders.
Appendix F Page F8

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 4 of 30)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Settlement
Pore Pressures in Core (Average Rockfill and/or Filters Comment Internal Monitoring
(from end of Steady / Instrument
Comments Comment mm/year) Comment Equipment
embankment Fluctuating Type/No.
mm % of H
construction)
IVMs in rockfill, upstream =
Filled in 5 months, but not sure of vertical strain profile normal, 0
During construction - high PWP in core, ru approx 1150 mm settlement or IVMs indicate deformation of rockfill
Agigawa - timing with respect to end of - - - IVM 2745 2.71% to 1% near crest increasing to -
0.5 (or greater). 1.35% and downstream = less than that of core.
construction. 3 to 4% near base.
905 mm or 1.48%.

Seasonal drawdown of 9 m (?). Only


Ajuare not known no information Fluctuating - - - - - - - - -
limited data.

Large internal deformations of the


Started Jan 1990. No indication of core during latter stages of
Ataturk -0.6 to 3.5 ? operation. Close to FSL by March - No indication of operation. - - IVM - - no information - construction / early FF as indicated -
1994. by loss of gauges at lower
elevations.

During construction - PWP up to 70% of applied


Bath County Upper Rapid daily drawdowns of up to 32 m load. Slow dissipation post construction. PWP
0.67 to 1.08 years Aug 1985 to Dec 1985 Fluctuating - - - - - - - -
Dam (for power generation) shows about 20% amplitude response to reservoir
level on rapid drawdown.

Partial impounding to within 23 m of IVMs indicate large


Deformation distorted IN tubes so no
FSL following heavy rain in early deformation of rockfill during
Steady for 5 years at FSL, then 17m Relatively high PWP in core during construction, Large plastic deformations in readings possible, indicating large
Beliche -0.2 to 4.1 years 1985, overtopping the cofferdam and Steady - IVM 2585 4.9 (1.5 to 7.4%) construction. Estimated IN, IVM in rockfill
drawdown. estimate r u = 0.3 to 0.6 at EOC. some zones of the core. internal lateral displacements of the
saturating the upstream rockfill. moduli for Zone 3A rockfill =
core.
Virtually full by Jan 1988. 5 to 27 MPa.

Generally steady at close to FSL.


only have records after 1984 (18 Only have post construction
Bellfield no information Steady Drawdowns up to 3 to 5m, but not - 927 IVM - - - - -
years after EOC) records after 1987
every year.

Generally within 3 to 4 m of FSL. During construction - positive PWP up to 35 to


Bjelki-Peterson 0.3 to 1.6 years Late 1988 to Mid 1990 Steady One drawdown of 12 m over several 130 kPa, ru up to 0.2 (but will be higher because 804 - - - - - - -
years. this does not allow for arching).

During construction - varied PWP response due to High vertical strains in wet
moisture variation. High PWP in the wetter placed placed region of the core (up to
HSGs indicate rockfill moduli is
earthfills (ru up to about 0.7), lower positive PWP 10 to 12%), lower strains in the
Annual drawdown, typically of 10 to highest in the inner Zone 3A region
in lower and upper drier placed core (r u up to 0.3 IVM, 1 (4.5m dry placed lower region of the
Started May 1968, reached FSL in 30 m. Larger drawdowns in 1979/80 (60 to 100 MPa) and least in the HSGs and IHMs in
Blowering 0.07 to 1.7 years Fluctuating to 0.4). Slow dissipation of high PWP over more - upstream of 5875 5.63 core (3 to 6%). Possible shear -
Nov-Dec 1969. of 40m, 1982/83 of 58m, 1987/88 of Zone 3B rockfill (10 to 20 MPa). downstream rockfill.
than 20 yrs. dam axis) development in the core during
35m, and 1997/98 of 54m. Therefore, large differential between
During operation - decreasing amplitude to the latter stages of
core and filters / inner Zone 3A.
reservoir fluctuations with increasing distance to construction at about 65 to 70
downstream. m depth.

During construction - negligible positive PWP in Considered as "normal"


May 1979 to Feb 1984. Fluctuated
Typically at FSL except for low period core. 1 IVM, 5 m behaviour. Strains range from
Chaffey 0.17 to 4.9 years. within this period before reaching Steady 800 911 1.51 - - -
from 1994 to 1996 (15 to 17 years). During operation - slow rise in PWP in core with u/str of axis. 0.5% near crest to 3% near
FSL.
time over 5 to 15 yrs. base (4.75% at contact).

May 1956 to July 1957. To within 15 Seasonal drawdown, 20 to 35 m in precipitation


Cherry Valley 0 to 1.72 years Fluctuating - - - - - - - -
m of FSL in 1956. first two drawdowns. mostly as snow.

Relatively large vertical strains


Started just prior to the end of Records only for first year post (6 to 7%) in region 85 to 105 m
Chicoasen 0 to 0.22 Steady - - IVM 6650 0.03 - - -
construction. construction. below crest, likely plastic
deformation.
Appendix F Page F9

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 5 of 30)
SURFACE MONITORING - POST CONSTRUCTION
Crest Upstream Slope Downstream Slope
Time of Initial
Name Readings Comments
SLT (% per Comments Comments on Deformation and Observations.
(years after Settlement Displacement Settlement Displace- Settlement Displace-
log cycle of
EOC) (mm) (mm) (mm) ment (mm) (mm) ment (mm)
time)

Only have deformation during construction. IVM indicates deformation of core


47 mm (on first Only post construction information is 47mm downstream
Agigawa no indication - - - - - - - shows normal vertical strain profile. High moduli of filters (285 MPa) and rockfill
filling) displacement of the crest on first filling.
(130 MPa).

Some but limited data. Not sure if monitoring captures the period of
138 mm or No indication of problems with embankment behaviour. Difficult to
Ajuare 0.25 0.085 310 mm (25 years) - - - - first filling. Relatively large downstream displacement for a dam of Relatively large downstream movement for embankment of this height.
0.30% (25 years) evaluate given limited amount of data.
this height.

Longitudinal cracks and scarps formed on the crest and


Very large magnitude of settlement of the crest during first filling.
slickensided surfaces found in the core near to its downstream
Latter period of monitoring indicates the settlement rate is very high. Case of 'abnormal' deformation behaviour. The magnitude of post construction
0 (crest) 7225 or 4.0% 924 mm or 1380 mm interface. Together with the large differential settlement between
Ataturk 3.05 - - - Much lower magnitudes of settlement of the downstream shoulder. crest settlement and settlement rate are clear outliers to similar embankment
0 to 1.5 (d/stream) (6.8 yrs) 0.86% (5.8 yrs) (5.8 yrs) the crest and downstream shoulder, a shear is likely to have
Large magnitude displacements to downstream of the downstream types. Suspect it is very likely that a shear has developed in the core.
developed in the core. Collapse compression of the upstream
shoulder.
rockfill being a significant factor.

Consider deformation as 'normal'. The deformation behaviour and The purpose of the dam is for power generation and it is subject to rapid daily
Bath County Upper 546 mm or Limited data. Crest settlement indicates 'normal' behaviour. Most of
0.17 0.37 190 mm (6.3 yrs) - - - - embankment performance would indicate the rockfills were drawdowns. Embankment constructed of weathered to fresh siltstones and
Dam 0.41% (6.3 yrs) downstream displacement occurred on first filling.
possibly well compacted. sandstones.

Large settlements on first filling, during the heavy rainfall in 1989 and A number of aspects of the deformation behaviour are considered
The rockfill was poorly placed, both upstream and downstream, and of poor
on drawdown in 1993. Internal deformations post construction 'abnormal' including the large magnitude settlement of the crest,
1038 or 1.89% 340 mm or 484 mm or 1.2% 300 mm quality adjacent to the core. It was susceptible to large deformations on collapse
Beliche 0 0.26 445 mm (9 yrs) - indicate large vertical strains in upper earth and rockfill zones. Large and acceleration of the settlement rate during heavy rainfall and on
(9 yrs) 0.74% (3 years) (9 yrs) (9 yrs) compression as the monitoring records indicate. Suspect the deformation
displacements to downstream on FF. On drawdown, the core drawdown. The large internal strains in upper core and upstream
behaviour is 'abnormal'.
displaced upstream whilst downstream shoulder remained steady. rockfill indicate plastic deformation or possible shearing.

0 to 0.3 (?) Missing period of SMPs records from base survey to 1987 (0/0.3 to Cracking on crest observed in 1985, but no details. Internal core
(assumed, no 496 mm or 285 mm or 457 mm 20 years after EOC). Long-term crest settlement rate on the high settlements indicate the presence of localised concentrated zones Overall, SMPs indicate 'normal' behaviour, although the long-term rates are on
Bellfield 0.50 298 mm (31 yrs) - -
indication of date of 0.95% (31 yrs) 0.72% (31 yrs) (31 yrs) side for steady reservoir condition. Displacement rate is also on the of vertical strain post construction. Those that developed after the high side.
installation) high side. 1987 are relatively minor and are not associated with drawdown.

Small magnitude deformations probably associated with low


0.41 years (started 40 mm or 0.10% 38 mm or 0.12% 15 mm 20 mm or 0.07% 11 mm Small magnitude deformations post construction. Very low long-term embankment height and well compacted condition of all
Bjelki-Peterson 0.045 7 mm (10 yrs) Consider deformation behaviour as "normal".
10 Feb 1989) (10 yrs) (10 yrs) (10 yrs) (10 yrs) (10 yrs) settlement rate. embankment zones. Embankment in a very good condition, no
cracking or settlement of crest.

Relatively large magnitude settlement of the crest and upstream The deformation behaviour shows several 'abnormal' trends. The
shoulder on first filling. Accelerations in settlement rate of the crest post construction IVM records show further shear type
785 mm or
and upstream shoulder on large drawdown in 1982/83. displacements at 65 to 70 m depth below crest level on first filling Monitoring records indicates likely shear development in the core that initially
0.73% (30 yrs)
0.24 years (0 for 895 mm or -430 mm 378 mm or 180 mm Displacements show the d/str shoulder displaced d/str on and after and possibly on large drawdown. On drawdown, the acceleration developed during the latter stages of construction and re-occurred during first
Blowering 1180 mm or 0.42 -85 mm (30 yrs)
IVM at crest) 0.84% (30 yrs) (30 yrs) 0.43% (30 yrs) (30 yrs) FF; limited displacement of the crest on FF and then a slight in settlement and non-recovered upstream displacements are filling and on subsequent large drawdown. The very large crest settlement on
1.12% from IVM
upstream trend thereafter; and the upstream shoulder shows indicative of shear deformation to upstream within the core. The first filling required the crest to be raised February 1969.
(28 yrs)
upstream displacement at an accelerating rate (in log time) and increasing rate of displacement (in log time) of the upstream slope
acceleration on large drawdown. is potentially "abnormal"

-19 mm (17.5 yrs), Up to 50 mm crest spreading over monitored period. Small


Small magnitude of displacements; upstream crest and slope
149 mm or u/str edge 113 mm or -38 mm 115 mm or 27 mm displacements possibly due to low FSL (15 m below crest level of
Chaffey 0.23 years 0.19 displaced upstream, downstream crest and slope displaced Consider as normal deformation behaviour.
0.25% (17.5 yrs) 28 mm (17.5 yrs), 0.22% (17.5 yrs) (17.5 yrs) 0.23% (17.5 yrs) (17.5 yrs) 58 m high embankment). Internal settlement profile in core is
downstream. Large proportion of settlement occurred on first filling.
d/str edge normal.

135 mm or Limited settlement of central region of the crest, most of which Differential settlement between core and rockfill resulted in
Limited settlement of the broad SM/ML core indicative of its high moduli in
0.08 years after occurred on first filling. Much greater settlement on FF of the cracking along the junction of the core and transition on both the
Cherry Valley 0.14% (2.5 0.03 106 mm (2.5 yrs) - - - - comparison to the dumped and sluiced rockfill. Deformation behaviour
EOC upstream rockfill. A large proportion of the crest displacement on a upstream and downstream edges of the crest. Collapse
years) considered "normal".
filling cycle is recovered on drawdown. compression on wetting contributed to settlement of the rockfill.

Indication of possible shear type movements in the upstream filter


849 mm or 682 mm or -43 mm On first filling concentrated zones of deformation in the upstream on first filling, but not in the core. Localised zones of high Dam constructed in a narrow steep sided valley. Significant arching in the
Chicoasen 0.1 - -43 mm (0.33 yrs) - -
0.33% (0.75 yrs) 0.27% (0.75 yrs) (0.3 yrs) rockfill and filters in the upper half of the embankment. deformation in rockfill possibly indicates localised zone of collapse narrow core, particularly in the deep cut-off trench.
compression.
Appendix F Page F10

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 6 of 30)
GENERAL DETAILS CONSTRUCTION / DESIGN
Embankment
Construction Timing Dimensions
Classification
Name Owner/ References
Location Geology Foundation Dam Core Upstream Downstream Comments on Design
Authority Year Time Core Height, Length, Ratio
Zoning Width / Slope Slope
Completed (years) n Type H (m) L (m) L/H
Class Slope (H to V) (H to V)

Central clayey sand core supported by reasonably to well


compacted rockfill (outer zone poorly compacted).
Australia, New State Water LWC NSW (1995a)
Copeton porpyritic biotite granite Rock 1973, June 1.75 5,2,0 c-tm SC 113 1484 13.1 1.7H to 1V 1.8H to 1V Cofferdam incorporated into the upstream toe. Embankment
South Wales DLWC PWD NSW (19XX)
slopes of 1.6H to 1V on the mid to upper slopes and 2H to
1V on lower slopes.

Central core supported by compacted rockfill shoulders.


folded quartzite, silicified CDW (1970)
Corin Australia, ACT ACTEW Rock 1968, Jan 1 5,2,0 c-tm SM 74 282 3.8 1.8H to 1V 1.8H to 1V Filters / transition 10.6 m wide downstream and 9 m wide
sandstone & siltstone. ACTEW (1994)
upstream of the core.

tuffaceous sediments - Very narrow core with slight upstream slope. Dual filters
US Corp of Rock - abutments and valley stripped to 3 Pope (1967)
Cougar USA, Oregon tuff and mudstone 1963, Oct 5,2,0 c-tn (u) GM 159 488 3.1 1.8H to 1V 1.6H to 1V downstream and single filter upstream of core. Dam axis
Engineers bedrock (seasonal) Cooke & Strassburger (1988)
Basalt on abutments arched to upstream.

Narrow core sloped slightly upstream of centreline supported Dolezalova & Leitner (1981)
by shoulders of compacted rockfill. Broad transition zone Brousek (1976)
Dalesice Checkoslovakia - - Rock 1977, Nov 6 5,2,0 c-tn (u) CL 90 330 3.7 1.65H to 1V 1.5H to 1V
downstream of the core. The outlet conduits fill the width of Vanicek (1982)
the narrow gully. Holomek (1994)

Central core of clayey and silty sands supported by SMEC (1975, 1997a)
Goulburn shoulders of dry placed and compacted rockfill. Dam located GMW (1999)
Dartmouth Australia, Victoria granitic gneiss Rock 1978, Nov 2.75 5,2,0 c-tm SC/SM 180 670 3.7 1.75H to 1V 1.75H to 1V
Murray Water in a valley with narrow river section and moderately steep Cole & Cummins (1981)
abutment slopes. Murley & Cummins (1982)

Thin slightly upstream sloping clay core supported by


Sowers et al (1993)
Sandstone and compacted to poorly compacted rockfill shoulders. Stability
Djatiluhur Farhi & Hamon (1967)
Djatiluhur Indonesia claystone, andesite on Rock 1965 3 (?) 5,2,0 c-tn (u) CH 105 1200 11.4 1.65H to 1 3.75H to 1 berm of random fill on the downstream slope due to
Authority Bister et al (1994)
abutments. slickensided surfaces in foundation. Upstream cofferdam
Sherard (1973)
incorporated into the upstream toe of the embankment.

Broad central core supported by poorly compacted rockfill


Sedimentary - Core on bedrock. Shoulders on either 2H to 1V (2.5H to shoulders of varying quality rockfill. Zoned core; inner zone SMEC (1999a)
Goulburn
Eildon Australia, Victoria sandstone, siltstone, bedrock or sandy gravels (1.5 to 3 m 1955, June 2 to 2.5 5,1,0 c-tk CL 79 937 11.9 2.5H to 1V 1V on lower (Zone 1) of clays and outer zone (Zone 1A) of clayey sands. Shaw (1953)
Murray Water
mudstone and shale depth). slopes) Core built up ahead of filters and rockfill during construction Speedie (1948)
(up to 18 m) .

Abutments on weathered bedrock. In


Very narrow central clay core (0.09H to 1V up and
CFE narrow river section core, filters and
Metamorhic - silicified downstream slopes) supported by shoulders of poorly Marsal & Ramirez de Arellano
(Federal inner downstream rockfill on bedrock;
El Infiernillo Mexico conglomerate, basaltic 1963, Dec 1.4 5,2,0 c-tn CL-CH 148 344 2.3 1.75H to 1V 1.75H to 1V compacted rockfill. Cofferdams incorporated into up and (1964, 1967, 1972)
Electricity upstream shoulder and outer
dykes. downstream shoulders. Constructed in narrow gorge with Alberro (1972)
Commission) downstream shoulder on river bed
steep abutment slopes.
deposits (boulders with fine sand).

Central clay core (0.5H to 1V downstream and 0.75H to 1V


Sedimentary - upstream) supported by dry placed and poorly compacted
Heitlinger et al (1965)
Goulburn sandstones and rockfill shoulders. Steep benched outer slopes; 1.3H to 1V
Eppalock Australia, Victoria Rock 1962, Mar 1.5 5,2,0 c-tk CL 47 700 14.9 1.9H to 1V 1.65H to 1V Woodward Clyde (1999)
Murray Water mudstones, dyke in lower slopes with 4.3 to 7.5m bench widths. Stabilising fill at lower
SMEC (1998d)
river section. upstream toe and cofferdam incorporated into upstream
rockfill toe.
Appendix F Page F11

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 7 of 30)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Relative
Layer
ASCS Ratio or Dry Moisture ASCS Density or Level of
Zone Source n Placement Methods Comment Zone Source n Thickness Plant Comment
Class Density Content Class Dry Density Compaction
(mm)
(t/m3 ) (t/m3 )

SC (some SM (Spec. >70% D R),


blend crushed Dual filters used up and downstream of core.
and CL) - clayey 100.1% SMDD Well compacted - 150 mm (Spec. 1% dry to 1% Clayey sand with medium plasticity fines. 2A & 2B 2A - SW/SP
Copeton 1 residual granite 3 granite & washed 1.85 (2A) & 2.08 375 to 450 minimum 1p 10t SDVR Good Combined width of 5 m downstream and 2.5 m
sand, 19 to 60% 1.76 t/m layers, sheepsfoot rollers wet) 1 m thick contact zone of CH/MH clays. (filters) 2B - GW/GP 3
sand (2B) t/m upstream.
fines

SM (SC) - Gravelly silty and clayey sands. Fines of Dual filters up (2A & 3A) and downstream (2B
Well compacted - 100 to 150 river sand &
slopewash & gravelly silty to 99.4% SMDD medium plasticity. Downstream core 2A, 2B & 3A and 3A). Zone 3A is rockfill transition of 6 m
Corin 1 3 mm layers, 12p (minimum) 0.3% dry of OMC gravels (2A, 2B) GW/GP (Spec. D R >70%) 450 10t SDVR, 4p for Zone 3A Good
weathered granite clayey sands, 1.75 t/m slope of 0.3H to 1V and upstream of 0.4H (filters) width. Combined width of 10.6 m downstream
sheepsfoot rollers rockfill (3A)
avg of 46% fines to 1V. and 9 m upstream.

GM - silty sandy Well compacted - 300 mm


weathered rock, 100% SMDD 2A & 2B alluvial/colluvial 3 45t pneumatic roller (2A) or Well-compacted filters. Slightly silty (3 to 20%
Cougar 1 gravel, 10 to 3 (loose) layers, 4p 45t 1% wet Well-compacted silty sandy gravel core. GP/GM 1.91 t/m 300 Good
talus and silt fines 1.89 t/m (filters) gravels and spalls 4p 10t SDVR (2B) fines).
30% fines pneumatic roller

Gravelly Sand Dual filter / transition zones up and downstream


(bulk density = Initially wet of OMC, Suspect well compacted given period of 2A & 2B (filter
Dalesice 1 colluvium CL 3 no information - (2A) & Sandy D R = 72% - no information - of the core. The downstream Zone 2B transition
2.09 t/m ) then at OMC construction. / transition)
Gravel (2B) is very broad (approx 36m base width).

SC/SM - clayey Clayey and silty sands. Fines of medium Well compacted filters of high moduli (as
(Spec. >98%
to silty sands, plasticity. Moisture spec. varied during 1 to 4p 14t SDVR. Started indicated by low deformation). Dual filters
SMDD) Well compacted - 150 mm 2A & 2B processed granitic 3
Dartmouth 1 residual granite avg 24% fines (1.0 dry to 2% wet) construction - initially wet of OMC, then GW/GP 2.2 t/m 500 at 4p, reduced due to high Good downstream (width increasing from 5 - 6 m to 12 -
1.82 to 1.91 layers, 8p sheepsfoot roller (filters) gneiss
(15 to 50% 3 dry of OMC for mid section, upper part moduli of filters. 13 m). Single 2A filter upstream of core (dual in
t/m
range) wet of OMC. crest region).

High plasticity clays and sandy clays


placed wet of OMC. Investigation in 1986
3
weathered CH - sandy 1.58 t/m , or average of 2.2% wet (Sowers er al 1993) found the mid to 2A & 2B water added during limited information. Dual filters used up and
Djatiluhur 1 no information - - - - -
claystone clays and clays 97% SMDD of OMC upper region of the core (above EL 65m) (filters) placement downstream of the core.
was much wetter and softer than
anticipated.

Very high compaction specification. Zone


Spec. > 100%
1 (clays of low to medium plasticity) used
Adjusted Proct Well compacted - 150 mm Spec 2% dry to 1% Zone 2 material dumped over edge of Zone 1.
alluvial and 1 - CL in central region of the core. Zone 1A 2 (filter / end dumped, no
Eildon 1 and 1A Test (3 layers, layers, 10p (minimum) wet of OMC (OMC for alluvial GP - - poor Width of 4.5 to 6 m. Fines content = 7%
colluvial 1A - SC/SM/GM (clayey sands with low to medium transition) compaction.
40 blows per sheepsfoot rollers adjusted test) average, but varies.
plasticity fines) used in the outer core
layer)
region.

Core of medium to high plasticity sandy Spec. D R >70% Dual filters placed up and downstream of core.
3 Reasonably to well 2A - alluvial sands
residual and CL/CH - sandy 1.59 t/m clays (54 to 70% fines) placed well wet of 2A & 2B 2A - SP reasonable to Width of filter zone increasing with depth from
El Infiernillo 1 compacted - 150 mm layers, 3.7% wet of OMC 2B - processed 1.87 (2A) to 2.00 300 4p 2t SDVR
alluvial (?) clays 93.7% SMDD OMC. Likely arching across very narrow (filters) 2B - GP 3 good approx. 2-5m near crest to 15-18m at base of
13.5t sheepsfoot rollers gravels (?) (2B) t/m
wet placed core. core.

2A - spread using spreader


Zone 2A and 2B used both up and downstream
CL - sandy Reasonably compacted - 380 Sandy clays of medium plasticity. Broad 2A - screened river box and compacted by light
98.8% SMDD 2A & 2B (filter 2A - GW/GC (2a - bulk density 2A - 450 2A - reasonable of the central core, width of 4.5 to approx. 7m.
Eppalock 1 alluvial clays, 50 to 85% 3 mm (loose) layers, 12p 0.8% dry of OMC central core zone of relatively large width gravel 3 track rolling
1.73 t/m / transition) 2B - GP = 1.94 t/m ) 2B - ? 2B - poor High clay fines content in upper part of Zone 2A
fines sheepsfoot rollers at crest (approx. 8.5m). 2B - crushed basalt 2B - end dumped in high
filter. Zone 2B is a rockfill transition zone.
lifts
Appendix F Page F12

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 8 of 30)
MATERIALS
Rockfill Zone Rockfill Zone (in some cases earthfill stabilising berm)
Name
Dry Layer Level of Dry Layer Level of
Grading Clean / Grading Clean /
Zone Source Density Thickness Placement Method Comp- Comment Zone Source Density Thickness Placement Method Comp- Comments
Description Dirty Description Dirty
(t/m 3) (m) action (t/m3 ) (m) action

Very high strength rock. Zone 3B used as


Very high strength rock. Zone 3A used 2p 10t SDVR, no water outer rockfill zone and also in bulk of the
3 4p 10t SDVR, no water reasonable to 3
Copeton 3A Granite - clean ? 1.99 t/m 1.2 adjacent to central core, both upstream and 3B Granite - clean ? 1.87 t/m 3.7 added. Also dumped in poor upstream cofferdam. It is located
added good
downstream. high lifts. immediately up and downstream of the filters
in the upper 25 to 28 m.

quartzite, Well-compacted rockfill. Zone 3B is inner quartzite,


Reasonable compaction of rockfill. Zone 3C
Corin 3B sandstone & - clean - 0.9 4p 10t SDVR Good rockfill shoulder zone adjacent to filters and 3C sandstone & - clean - 1.8 4p 10t SDVR Reasonable
is outer rockfill shoulder zone.
siltstone transition. siltstone

sound (3B) to Poor to reasonably compacted rockfill used


Sound basalt and 300 mm max Well compacted Zone 3A rockfill used in the 3B - max 600 mm
3 2p 45t pneumatic or 4p weathered (3C) 3B - clean 3B - 0.9 m 3B & 3C - 2p D8 dozer, no Poor to in the upstream shoulder and outer d/str
Cougar 3A andesite (H to no fines clean 1.82 t/m 0.45 to 0.6 Good in the inner downstream shoulder. Moduli 3B & 3C 3C - up to 25% < -
10t SDVR basalt and 3C - dirty 3C - 0.6 m indication if water added reasonable shoulder. Lesser quality Zone 3C used in the
VH) Cu ~ 10 approx. 40 to 60 MPa. 4.75 mm
andesite inner upstream shoulder.

(bulk density = Rockfill indicated as being of poor quality.


Dalesice 3 - - - 3 1.0 to 1.5 6p 8.3t & 13t SDVR Good - - - - - - - - -
2.40 t/m ) Well compacted.

Well compacted rockfill, placed without


1.0 (0.5 m Zone 3B used in outer rockfill shoulders, both
granitic gneiss max 600 mm. 15% 3 4p 14t SDVR, no water water addition. Zone 3A used upstream granitic gneiss max 1.5 m clean ? 3 4p 14t SDVR, no warer
Dartmouth 3A clean 2.1 t/m within 6 m of Good 3B 2.1 t/m 2.0 reasonable upstream and downstream. Placed without
(SW to Fr) pass 19 mm. added and downstream as inner shoulder rockfill (SW to Fr) (coarser than 3A) added
filters) water addition.
adjacent to the central core.

Not precisely known.


Roller compacted and no
formal compaction Limited details. Farhi and Hamon (1967) Random - all Random fill zone acting as stabilising berm
compacted, no details on
Djatiluhur 3 andesite - - - 0.5 to 2.0 (trafficked by trucks and poor (?) indicate that the rockfill was not formally 4 types of materials - - - - - on downstream slope due to presence of
methods.
bulldozers). Water compacted in the upper elevations. used slickensides in foundation.
added, from 30 % to
300% by volume.

Zone 3B is poorest quality rockfill (random


Tipped over advancing Zone 3A is 'first' quality rockfill described on dirty (prone to Tipped (over advancing rockfill) and was used in the outer
quartzitic face and spread by sections. It was used in the outer rockfill sedimentary rock breakdown, face) and spread by downstream shoulder. Zone 3C descibed as
Eildon 3A gap graded clean ? - 2.0 Poor 3B & 3C - - 2.0 Poor
sandstone dozers. No indication zone of the upstream shoulder and inner types high fines dozers. No indication that medium quality rock and was used in the
that water was added. rockfill zone of the downstream shoulder. content) water was added. inner upstream shoulder, and contained
gravel drainage layers.

Rock of very high strength (UCS = 125


Rock of very high strength (UCS = 125 MPa)
Diorite and 3 MPa) not susceptible to strength loss on Diorite and 3
Max size 600mm 1.85 t/m 1.76 t/m not susceptible to strength loss on wetting.
silicified 4p D8 dozer, no water wetting. Zone 3A used as inner rockfill silicified Max size > 600 No formal compaction and
El Infiernillo 3A 23% < 20mm clean to dirty 32% voids 0.6 to 1.0 poor 3B - 35% voids 2.0 to 2.5 poor Zone 3B used as outer rockfill zones in up
conglomerate, added shoulder zones up and downstream of the conglomerate, mm no water added.
2% fines ratio ratio and downstream shoulder. Moduli estimated
sound central core. Moduli estimated at 30 to 50 sound
at 15 to 30 MPa.
MPa.

2.0 to 4.0, High to very high strength basalt rock. 4 sandstone and spread by dozer and
Max size ~ 1.0m bulk density = Spread by tractor, no Limited information available. Used in the
Eppalock 3 basalt dirty 3 one 10m poor Zone 3 used in up and downstream (stabilising shale (HW to - dirty - - compacted by Poor
20 to 25% < 25mm. 2.05 t/m water added stability berm on the upstream slope.
dumped layer shoulders. fill) MW) construction traffic
Appendix F Page F13

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 9 of 30)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Settlement
Pore Pressures in Core (Average Rockfill and/or Filters Comment Internal Monitoring
(from end of Steady / Instrument
Comments Comment mm/year) Comment Equipment
embankment Fluctuating Type/No.
mm % of H
construction)

During construction - positive PWP in contact clay


Some differential settlement during
Variable water level, but not annual only (r u approx 0.5), dissipating over 15 to 20 yrs. Strain profile look normal
Impounded water before EOC, construction between core and HSGs in d/stream filter &
drawdown and generally not a rapid During operation - core responds to reservoir 655 to 3575 (increasing strain with
Copeton -0.8 to 5.2 years reached within 45m of FSL before Fluctuating - IVM, 4 in core 1.27% to 3.34% - rockfill (greater settlement of core). rockfill. IHMs - 3 rows in
drawdown; extended low periods fluctuation, but with reducing amplitude and mm increasing depth). High strains
EOC. Reached FSL in August 1978. Possibly some arching, but core main section.
followed by rapid rise. greater delay with increasing distance from the near base in contact zone.
relatively broad (0.4H to 1V slopes).
upstream face.

seasonal drawdown of less than 10 m


Started in May 1968. Not sure when
Corin 0.31 to ? Steady from 1984 to 1988, since then has - - - - - - - - -
FSL reached.
been steady

During construction - positive PWP, ru > 0.2 to 0.5.


0 to 0.67 years (Nov Reservoir at 57 m below FSL when Rapid dissipation of construction PWP after EOC. Zone 3A rockfill - 2718 mm Deformation of Zone 3A indicates
Cougar Fluctuating seasonal drawdown of 40 to 70m - - - - - -
1963 to July 1964) monitoring started. PWP respond to reservoir level, but at reduced (up to 3%) moduli = 40 to 60 MPa at EOC.
amplitude.

During construction - high PWP in lower, wetter


July 76 to Jan 80. First filling started placed region of the core (r u ~ 0.96). Of lesser
Dalesice -1.3 to 2.2 years prior to EOC. Within 9 m of FSL at - no information magnitude in the mid to upper region (r u = 0.3 to - - - - - - - -
EOC. 0.67). EPWP still present after 15 years post
construction.

During construction - High PWP in lower region of


3220 (ES2 Variable strain profile (0.5% to
Started Nov 1977, 1 year prior to core due to placement wet of OMC (ru up to 0.85). HSGs indicate filters and transition
Not an annual drawdown. Significant H=104m) 6 - 7%). Suspect due to HSGs in core, filter &
EOC. Reached within 70 m of FSL Reduced moisture spec for mid sections (ru = 0.45 3.10% (ES2) and (Zone 3A) of high moduli. Large
Dartmouth -1 to 11 years Fluctuating drawdowns in 1983 (of 70m), 1994 (of - IVM, 2 in core continued adjustment of - downstream rockfill; IHMs in
prior to EOC. Reached FSL in Nov to 0.65). 3.95% (ES1) differential deformation between core
20m) and 1997/98 (of 50m). 7070 (ES1 moisture spec. during downstream shoulder.
1989 (within 20 m at 3.7 years). Post construction slow dissipation of construction and filters, and filters and rockfill.
H=179m) construction.
PWP.

Large deformations of the crest and


Annual seasonal drawdown of 10 to upstream slope during latter stages
1986 investigation indicated high pore pressures
-0.75 to between 2 - 2.5 Water impounded during construction 18 m. Largest drawdowns in 1972 of construction. Crest cracking
Djatiluhur Fluctuating still present in core (20 years after construction), - - - - - - SMPs
years (to EL 80m or 31 m below FSL) (27m to EL 78m) and 1982 (30m to (longitudinal) near the downstream
possibly affected by drilling and installation.
EL 77m). core interface observed during a
shutdown period toward the EOC.

During construction - only minor PWP developed


Impounding began prior to the end of Annual drawdown, typically 5 to 15 m. due to dry placement of core (dry of Standard
construction and reached to within Larger drawdown events of 18 to 28 optimum). IVMs installed in core, but no results
Eildon -0.85 to 1.2 years Fluctuating - - - - - - -
approx 30 m of FSL at EOC. m in 1967/68, 1981-83 (28 m), During operation - upstream Zone 1A responds during construction available.
Reached FSL in July 1956. 1994/95, and 1997/98 (27m). with reservoir fluctuation. Large drop in head
across inner core (Zone 1).

Records indicate 'normal' IVM D1 (Zone 3A) vertical Vertical strain profile in Zone 3A
vertical strain profile, 0 to 1% strains of < 1% in top 30m rockfill similar to core. Vertical
INs, IVMs in core, filters and
El Infiernillo 0.5 to 0.85 years June to October 1964 Fluctuating Annual drawdown of 10 to 30m - - IVM, I1 3815 3.00% in top 20m increasing to 4 to increasing to 4 to 6% below strains in outer dumped rockfill
rockfill.
6% below 80m. Likely affected 80m depth. Higher vertical (Zone 3B) higher than core and Zone
by arching across narrow core. strains in Zone 3B (IVM I3). 3A.

Annual drawdown, typically 3 - 5m. During construction - positive PWP only observed Vertical strain profile at EOC
Larger drawdown events of 7 to 9.5 m in some piezometers in the core, maximum ru of shows normal behaviour; 0 to
Started May 1962, reached FSL inclinometers installed 1997
Eppalock 0.18 to 1.65 years Fluctuating in - 1967/68, 1976/78, 1982/83 0.22 (DCP1). Slow dissipation of PWP. - IVM, 1 1015 2.28% 0.5% in upper 20m decreasing - -
Nov/Dec 1963. to 1999
(8.5m), 1994/95 (9.5m) and 1997/98 During operation - delayed and reduced amplitude to 1.6 to 1.8% at depth (35 to
(8.5m). of response relative to the reservoir level. 45m).
Appendix F Page F14

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 10 of 30)
SURFACE MONITORING - POST CONSTRUCTION
Crest Upstream Slope Downstream Slope
Time of Initial
Name Readings Comments
SLT (% per Comments Comments on Deformation and Observations.
(years after Settlement Displacement Settlement Displace- Settlement Displace-
log cycle of
EOC) (mm) (mm) (mm) ment (mm) (mm) ment (mm)
time)

Displacement of all SMPs initially downstream on FF. Post FF crest Greater settlement of the upstream side of the crest is visually Possible shear development and continued deformation in this zone is possibly
781 mm or
-8 mm (25 yrs), and upstream slope trend to upstream, whilst downstream slope evident, and transverse and longitudunal cracks were observed in indicate of "abnormal" deformation behaviour. Timing of shear deformation is on
0.72% (25 yrs),
u/str edge 817 mm or 61 mm 501 mm or 238 mm continues downstream. Greater settlements of upstream crest and the bitumen surfacing on the crest. Suspect it is partly due to first filling and thereafter during rising reservoir conditions. Acceleration in
Copeton 0 185 mm 0.35
70 mm (25 yrs), 0.79% (25 yrs) (25 yrs) 0.53% (25 yrs) (25 yrs) slope. Internal settlement profiles in core indicate large strains at 20 collapse settlement of the outer dumped rockfill. Large strains settlement of the upstream slope and upstream crest during these periods
differential to
d/str edge m depth in IVM C and 25 m depth in IVM A, both on upstream slide measured in upstream IVM indicate possible shearing in the upper suggest differential movements triggering shear displacement. Otherwise trends
upstream
of dam axis. core near to its upstream interface. in SMPs appear to be normal.

Gap of 7 years from base survey to second reading. Very low long- Consider as 'normal' deformation behaviour. Query the base
83 or 0.11% (26 189 or 0.27% 39 mm 145 or 0.21% 106 mm Consider as 'normal' deformation behaviour. Well compacted central core of
Corin 0.29 years 0.04 59 mm (26 years) term settlement rate. Displacement is initially downstream for the survey of some SMPs (i.e. base date is possibly before EOC for
years) (26 years) (26 years) (26 years) (26 years) gravelly silty sand supported by compacted rockfill shoulders.
upper slopes and crest, and then shows a slow upstream trend. SMPs on the shoulders).

Upstream crest deformation suggests collapse compression of


292 to 547 mm (3.1 Upstream rockfill adjacent to filters comprised reasonable to poorly compacted
0.12 years after Acceleration in upstream crest settlement on first filling and during upstream rockfill on FF. Cracking on crest at both the up and
806 or 0.59% years), upstream weathered rockfill. Suspect this material was prone to collapse settlement as
Cougar EOC (started 0.13 - - - - both first and second drawdowns. Displacement records indicate downstream core interface, both with vertical differential settlement
(3.1 years) and downstream well as outer Zone 3B to some degree. Deformation of upstream crest and
9/12/63) crest spreading. Long-term settlement rate is normal. across the crest to upstream. Possible shear development in the
edge cracking indicative of "abnormal" deformation on and shortly after first filling.
core on drawdown.

no indication First filling to within 9 m of FSL prior to end of construction will


389 or 0.46% 324 or 0.39% 43 mm 131 or 0.18% 125 mm Limited data records. Upper slope and crest regions all moved Limited deformation records. Unusual design in that outlet conduits fill the width
Dalesice (assume close to - - have affected the magnitude of the post construction
(15 years) (15 years) (15 years) (15 years) (15 years) downstream on and after first filling. of the river gully. Embankment then constructed above this.
EOC) displacements measured from SMPs.

Large settlements during early stages of FF. Greater settlement of Acceleration in settlement on drawdown indicates potentially
upstream shoulder, probably due to collapse compression in the dry 'abnormal' deformation behaviour. On drawdown in 1983 the crest
1189 mm or 1262 mm or 138 mm 844 mm or 282 mm placed upstream rockfill on wetting. Acceleration in the settlement and upstream shoulder settled 100 to 150 mm, and in 1997 settled Well instrumented dam. Increase in the rate of settlement of the crest and
Dartmouth 0.05 years 0.37 29 mm (20 yrs)
0.67% (20 yrs) 0.88% (20 yrs) (20 yrs) 0.59% (20 yrs) (20 yrs) rate of the crest and upstream slope during drawdown in 1983 and 30 to 40 mm. Difficult to evaluate IVM data due to the moisture upstream slope on large drawdown in 1983.
again on drawdown in 1997. Large fluctuations in crest displacement content variation of the core, but possible plastic deformations in
with fluctuations in reservoir. the wetter placed core regions.

Longitudinal cracking in crest observed during construction.


Large magnitude settlements on first filling. Long-term settlement Sowers et al (1993) indicate limiting stability of the upstream shoulder above EL
- 104 Further longitudinal cracking soon after EOC. Pits in the core
2480 or 2.36% rate for the crest is very high. Acceleration in crest settlement on 65m (upper 50 m), presumably under a drawn down condition. Drawdown
0 years (at EOC on (1.2 yrs), 476 mm encountered horizontal cracks in the downstream region (Sherard
Djatiluhur (in 27 yrs) in 1.17 - 509 (0.6 yrs) 312 (1.2 yrs) large drawdown in 1972 and 1982, similar to increasing magnitudes limitations involked. Similar to increasing magnitudes of crest settlement on
1 Sept 1965). moved (1.2 yrs) 1973). Consider the deformation behaviour as "abnormal". In
centre. of settlement on each drawdown. Large downstream displacement similar magnitude large drawdowns is an unusual observation and potentially
upstream particular the acceleration in settlement on drawdown and high
of the crest and downstream shoulder on FF. indicative of the limiting stability condition of the upstream shoulder.
long-term settlement rate.

Large magnitude settlements of the up and downstream shoulders.


0.19 years - SMPs 525 mm or Several aspects indicative of potentially "abnormal" behaviour
-35 mm (13.4 to 43 Suspect it is largely due to collapse compression on wetting of the
on slopes. 0.66% (in 43 1175 mm or 1475 mm or including acceleration in settlement on drawdown. IVM records
yrs). Does not 217 mm 1328 mm dry placed and poorly compacted rockfill. Large downstream Potentially "abnormal" aspects of the post construction deformation behaviour.
Eildon 2.5 yrs - crest sett, yrs). Does not 0.52 1.69% (in 43 2.18% (in 43 indicate localised zones of high vertical strain in the core post
include first 13.4 (43 yrs). (43 yrs). displacement of downstream shoulder over first 5 years after EOC. Possible shear development in the core as indicated from IVM records.
13.4 yrs - crest include first 2.5 yrs). yrs). construction, with deformation in these zones on large drawdown.
yrs. Acceleration in settlement of the upstream crest and shoulder on
displ. yrs. This indicates possible shear development within the core.
large drawdown in 1968 and 1981/83.

Overall, the deformation behaviour is considered "normal", but it


Large magnitude of settlement of the crest and upstream slope on
580 mm or does show several "abnormal" trends. Collapse settlement of the
0.18 yrs (sett FF. Displacement of the crest is upstream in early stages of FF and Well instrumented embankment. Transverse cracking observed on the upper
1190 mm or 0.52% (in 8 yrs), rockfill on wetting contributed to the large settlements of the
SMPs) 930 mm or 405 mm then downstream in latter stages, thereafter it is downstream. From abutments, first noticed 3 days after the start of reservoir impounding. Possible
El Infiernillo 0.81% (in 17 0.09 348 mm (17 yrs). - greater for SMPs shoulders. The post FF acceleration of the settlement and
0.45 to 0.54 yrs 0.83% (in 8 yrs). (8 yrs). late 1966 to 1970 (all SMPs), sustained period of acceleration in shear development in the core post first filling, possibly drawdown related, but
yrs). over Zone 3B displacement was coincident with periods of heavy rainfall and tail-
(displ SMPs) settlement and downstream displacement occurred. Acceleration in not sure.
(0.66% in 8yrs) water inundation of the downstream toe. IVM data indicates
crest settlement on earthquake (7.6 mag.) in 1979.
possible shear development in the core.

Large settlements of upstream shoulders on FF and downstream


Consider as case of "abnormal" deformation behaviour. Shear
shoulder in first 4 years, indicative of collapse settlement on wetting.
development in the core confirmed from inclinometers and Case study of "abnormal" deformation behaviour. Woodward Clyde (1999)
Acceleration in settlement rate of SMPs on the crest and upstream
195 mm (CS2, observation during remedial works. Collapse compression of the indicated the factor of safety of the upper region of the upstream shoulder was
510 mm or 570 mm or -240 mm 840 mm or 750 mm shoulder on large drawdowns from 1982/83 on, with similar to
Eppalock 0.26 years 0.91 37yrs), -22 mm rockfill is a significant factors contributing to the shear approaching a marginal condition. The similar magnitude of crest settlement on
1.11% (37yrs) 1.60% (37yrs) (37yrs) 2.38% (37yrs) (37yrs) increasing magnitudes at CS1 for subsequent drawdowns. Crest
(CS1) development due to the limited lateral support afforded to the core large drawdowns of similar magnitude in unusual. Remedial works were
displacement shows change in trend from downstream to upstream,
from the shoulders. Longitudinal cracking in the crest was first undertaken in 1999 to address the dam safety issues.
and also non-recovered upstream displacement on the 1994/95
observed about 1973 and has since been persistent.
drawdown.
Appendix F Page F15

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 11 of 30)
GENERAL DETAILS CONSTRUCTION / DESIGN
Embankment
Construction Timing Dimensions
Classification
Name Owner/ References
Location Geology Foundation Dam Core Upstream Downstream Comments on Design
Authority Year Time Core Height, Length, Ratio
Zoning Width / Slope Slope
Completed (years) n Type H (m) L (m) L/H
Class Slope (H to V) (H to V)

Narrow central core with central diaphragm wall. Dual filters


Core on unweathered gneiss, shoulders
Frauenau Germany - Gneiss 1981, Aug ? 5 5,2,0 c-tn SM 80 670 8.4 1.5H to 1V 1.5H to 1V up to 10 m wide up and downstream of the core. Dam axis List & Beier (1985)
on weathered gneiss.
arched upstream.

Sedimentary - sandstone Central core with shoulders of weathered rockfill. Relatively


Fukada Japan - Rock 1976, Oct 1.1 4,1,2 c-tm GC (?) 55.5 340 6.1 2.75H to 1V 2.75H to 1V Yasunaka et al (1985)
and mudstone flat shoulders for embankment of this type and height.

1.75H to 1V Pinkerton & Paton (1968)


foiliated granodiorite and Central narrow core with inner shoulders of compacted
Australia, New 2.5 (2H to 1V from Hilton et al (1974)
Geehi SMHEA diorite (partly Rock 1966, Jan 5,2,0 c-tn SM 91 265 2.9 1.75H to 1V rockfill and outer shoulders of dumped rockfill. Dual filters up
South Wales (seasonal) 45m below Grimston (1989)
metamorphosed) and downstream.
crest level) GHD (1995a)

Core on bedrock.
Narrow central core with broad filter/transition zones and Lauffer & Schober (1964)
Foliated gneiss, glacial Shoulders on over-consolidated 3.5
Gepatsch Austria TIWAG 1964, Nov 5,2,0 c-tn GM/GC 153 605 4.0 1.5H to 1V 1.5H to 1V shoulders of reasonably compacted rockfills placed without Schober (1967)
activity alluvium and colluvium up to 50 m (seasonal)
water addition. Slight upstream curvature to dam axis. Schwab (1979)
deep.

Central core supported by narrow width filters and rockfill


Sedimentary -
Glenbawn, Saddle Australia, New State Water shoulders. Former spillway to main dam at upstream toe. Volk (1987)
mudstones with igneous Rock 1986, Aug 0.5 5,2,0 c-tm - 35 585 16.7 1.5H to 1V 1.5H to 1V
Dam A South Wales DLWC Saddle dam constructed as part of the works associated with LWC NSW (1995b)
dyke (diorite) intrusions.
raising of the main embankment.

Stafford & Weatherburn (1958)


Sedimentary - 3.4H to 1V (3H 3H to 1V (2.5H to
Broad central zoned core supported by broad transition zone Wilson & Scott (1957)
Glenbawn - main dam Australia, New State Water sandstone, propylites, to 1V upper 1V upper slope
Rock - 'soft and weak' 1957, Sept 3.5 5,1,0 c-tk CL 76.5 823 10.8 (up to 40 m width near base of dam) and poorly compacted WCIC (1982)
(prior to raising) South Wales DLWC breccia & tuff; igneous slope reducing reducing to 5H to
rockfill shoulders. Cofferdam at upstream toe. Volk (1987)
dykes and sills. to 4H to 1V) 1V)
LWC NSW (1995b)

dacite and meta-


62 (before Goldsmith (1977)
Googong Australia, ACT ACTEW sediment intruded by Rock 1977, Apr 1 5,2,0 c-tm SC 423 6.8 1.8H to 1V 1.7H to 1V Central core supported by compacted rockfill shoulders.
raising) ACTEW (1993)
granite.

Sedimentary / Medium width central core (0.28H to 1V slopes) supported by


Kansai Electric Kisa & Fukuroi (1994)
Kisenyama Japan Mertamorphic - slate, Rock 1963 unk 5,2,0 c-tm unk 88 255 2.9 2.5H to 1V 2.2H to 1V rockfill shoulders. Filters increasing in width with depth.
Power Co. Nose (1969)
sandstone & chert Relatively flat upstream slope.

The narrow core is aligned slightly to upstream (downstream


Kisa & Fukuroi (1994)
Kurokawa Japan - Tuff and breccia Rock 1973, Nov 2.5 5,2,0 c-tn - 98 325 3.3 2.5H to 1V 1.85H to 1V slope of core is near vertical). Filters increasing in width with
JNCOLD (1976)
depth. Relatively flat upstream slope.

Narrow core aligned slightly to upstream. Supported by Pare et al (1982)


Granite. Over-burden of compacted rockfill shoulders, with dual filter zones up and Pare (1984)
La Grande 2 (LG2) Canada, James Bay Hydro Quebec mainly on bedrock. 1978, Oct - 5,2,0 c-tn (u) SM 160 2900 18.1 1.8H to 1V 1.6H to 1V
glacial origin. downstream of the core. Cofferdam incorporated into the ICOLD (1989)
upstream shoulder. Dascal (1987)

Sedimentary - clay
Nutt (1975)
shales, siltstones, Weighting berms added both upstream and downstream
Australia, QDNR (now Rock. Overburden and weathered QWRC (19XXa)
Maroon sandstones. Basalt sills 1973, June 1.7 5,2,2 c-tk SC/CL 52 460 8.8 1.5H to 1V 1.5H to 1V during construction due to presence of pre-sheared clays
Queensland Sun Water) shale removed. Sbeghen (1990)
intruded along bedding seams in foundation.
Coffey & Hollingsworth (1971)
seams.

Abutments on ignimbrite. In broad


Central (zoned) core aligned slightly to upstream (slopes of
valley, core on heavily over- Penman (1988)
Volcanic (ignimbrite) over vertical downstream and 0.6H to 1V upstream) supported by
Matahina New Zealand Trust Power consolidated Tertiary sediments (clays, 1966, Oct (?) 2 5,1,0 c-tm CL/ML 85 400 4.7 2.5H to 1V 2.6H to 1V Galloway (1970)
tertiary sediments poorly compacted rockfill shoulders. Constructed in a wide
gravels and sands) and shoulders on WGL (1994)
valley with steep abutment slopes (approx. 50 degrees).
alluvial gravels
Appendix F Page F16

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 12 of 30)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Relative
Layer
ASCS Ratio or Dry Moisture ASCS Density or Level of
Zone Source Placement Methods Comment Zone Source n Thickness Plant Comment
Class
n Density Content Class Dry Density Compaction
3 3 (mm)
(t/m ) (t/m )

Silty sands with low fines content and 2A & 2B no information. Dual filters of narrow width up
Frauenau 1 residual soil SM - silty sands - no information - - - - - - -
small clay fraction. (filters) and downstream of the core.

weathered GC (?) - mix Limited information. Suspect reasonably No information. Single filter up and d/stream of
Fukada 1 sandstone and clay, sand & - no information - to well compacted given practices used in 2 (filter) - - - - - - the core. Dual horizontal filter in the lower region
granite gravels Japan at this time. of the d/stream shoulder.

Silty sand with fines (43% average fines Well compacted filter zones up and downstream
Well compacted - 150 mm
completely SM - silty sands, content) of low plasticity. Difficulty with 2A & 2B 2A - SP/SW (Spec. DR >70%) 2p 8t SDVR & 1p 20t
Geehi 1 100.6% SMDD layers, 12p 3.5t sheepsfoot 0.5% wet of OMC crushed rockfill 300 Good of the central core, of thin width (combined width
weathered gniess avg 43% fines moisture content due to breakdown of (filters) 2B - GP/GW easily met crawler tractor
rollers. = 7.3 m).
earthfill on compaction.

GM/GC core placed wet of OMC. Reasonable to good compaction as indicated by


3
GM/GC - silty to Well compacted - 300 mm 2% wet Upstream Zone 1A region of core had 1% 2.25 t/m 2 - trafficked only 2A - well
talus (colluvium) 96 - 99% SMDD 2 & 2A (filter / the low compressibility of these layers. Zone 2
Gepatsch 1 & 1A clayey fravels, 3 layers, 6p 40 t rubber tyred (0.5 - 1% wet from bentonite added. After May 1663 river deposits GP (n = 15.2%, e = 600 2A - 4p 40t rubber tyred compacted, 2 -
and morraine 2.18 t/m transition) used up and downstream of the core, Zone 2A
17 to 42% fines roller 5/63) improved compaction due to lower 0.18) roller or 4p 8.5t SDVR reasonable
mainly in downstream shoulder.
moisture content.

Limited information. Suspect reasonably


Glenbawn, Saddle to well compacted. Sloped at 0.45H to 2A & 2B Limited information. Dual filters used up and
1 - - - no information - crushed rock (2B) - - - - -
Dam A 1V upstream and 0.15H to 1V (filters) downstream of the core.
downstream.

Zone 1A, medium to high plasticity sandy Zone 2 transition up and downstream of the
1A - CL some Well compacted - 200 mm clays, is the inner core region. Zone 1, core. Of high relative density achieved by track
Glenbawn - main dam 2 (filter / river gravel & trafficking by dozers and reasonable to
1A & 1 alluvial / colluvial ? CH > 100% SMDD (loose) layers, ballasted 2.0% dry of OMC clayey sands with low plasticity fines, is GC/GP (?) DR ~ 82% 300 rolling. Contains up to 20% fines. Of broad
(prior to raising) transition) shingles trucks good
1 - SC tamping rollers the outer core region up and downstream width, from several metres wide at the crest to
of Zone 1A. about 35 m wide near the base.

Well compacted - 165 mm Clayey and silty sands to sandy clays, low Zone 2 & 3A used up and downstream of the
SC (some CL (Spec. >98% (Spec. 1% dry to 1% 2 & 3A (filters / river gravels & 3 2p 10t SDVR (2)
Googong 1 colluvial layers, padfoot rollers (Cat plasticity. Fines content = 44% avg (30 to transition) - 2.10 t/m (2) 500 Good central core. Zone 3A rockfill is a transition of ~
and SM) SMDD) wet) crushed rock 4p 10t SDVR (3A)
835) 55%). Core aligned slightly to upstream. 6 m width.

2A and 2B No information. Dual filters used up and


Kisenyama 1 - unk - no information - No information. - unk - - - -
(filters) downstream of the central core.

2A and 2B No information. Dual filters used up and


Kurokawa 1 - unk - no information - No information. - unk - - - -
(filters) downstream of the central core.

Well compacted gravelly silty sands (max


SM - gravelly Well compacted filters. Dual filters used up and
Well compacted - 450 mm size 250 mm, non-plastic fines). Core 2A & 2B 2A - gravels (Spec. 90% 3-4p 3-5t SDVR, water downstream of the central core. Width
silty sands, avg 2.07 to 2.11 Spec. 1% dry to 2%
La Grande 2 (LG2) 1 glacial till 3 layers, 4p 45t pneumatic aligned slightly upstream (upstream slope 2A - SP/SW 450 Good
29% fines (non- t/m wet of OMC (filters) 2B - crushed stone results >70% DR ) added if necessary increasing with depth, up to 20m width upstream
rollers. of 0.44H to 1V and downstream of -0.09H
plastic) and 25 to 30m width downstream of core.
to 1V).

SC/CL - sandy Filter / transition zones used up and downstream


(Spec. >98% Well compacted - 150 mm Spec. 1% dry to 2% Well compacted sandy clays and clayey 4p 4.5t (min) SDVR, wetted
Maroon 1 alluvial / colluvial clays and clayey 2A, 2B (filters) sands and gravels sands to gravels (Spec. >70% DR ) 300 Good of the core, and on the foundation under the up
SMDD) layers, tamping rollers wet of OMC sands of medium plasticity. prior to compaction
sands and downstream shoulders.

Lower region of the core placed well dry


Limited information. Considerable grading
greywacke, dry of OMC, lower of OMC; this region was of very stiff
CL - gravelly Well compacted - 150 mm Ignimbrite, GW/GM - sandy variation. Used up and downstream of central
Matahina 1 extremely 97% SMDD (?) section 1% to 1.5% consistency and brittle. Zone 1 is the 2 (transition) - - - -
clay layers, 16p grid rollers (?) weathered to silty gravels core (Zones 1 and 1A). Of broad width in lower
weathered dry. narrow central region of the core encased
section.
in Zone 1A.
Appendix F Page F17

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 13 of 30)
MATERIALS
Rockfill Zone Rockfill Zone (in some cases earthfill stabilising berm)
Name
Dry Layer Level of Dry Layer Level of
Grading Clean / Grading Clean /
Zone Source Density Thickness Placement Method Comp- Comment Zone Source Density Thickness Placement Method Comp- Comments
Description Dirty 3 Description Dirty 3
(t/m ) (m) action (t/m ) (m) action

Frauenau 3 - - - - - - - No information - - - - - - - - -

Weathered No information, but suspect compacted


presume
Fukada 3 sandstone and - - - - - given Erc estimated at 42 MPa from IVM - - - - - - - - -
dirty
granite. data.

Rock described as 'hard, durable and Rock described as 'hard, durable and
Max. 600 mm 1.8 to 30
3 2p 8t SDVR & 2p 20t reasonable to massive". Zone 3A is the inner rockfill zone Max size of 2.7 m massive". Zone 3B is outer rockfill zone up
Geehi 3A granitic gneiss 6.3% finer than 19 clean 2.00 t/m 0.9 3B granitic gneiss clean - (30 m in lower Dumped & sluiced Poor
crawler tractor, sluiced good located up and downstream of the central (coarse rockfill) and downstream. 3 m layer thickness in
mm. sections)
core (outer slope = 0.75H to 1V). upper region.

Finer material placed toward central


3
gneiss and Coarse (max 1.7 m) 1.89 t/m section, coarser toward outer slopes.
clean to dirty 4p 8.5t SDVR, no water
Gepatsch 3 oversize (> 200 8 - 45% < 20 mm (n = 28.8%, e 2.0 Reasonable Described as "firm" haer eye gneiss. Thin - - - - - - - - -
(variable) added
mm) from Zone 1. < 8% fines. = 0.41) Zone 3A transition between filters and
rockfill.

Suspect compacted given high density.


Glenbawn, Saddle limestone and 3
3 - - 2.12 t/m - limited information - Rock of very high strength. Zone 3 used in - - - - - - - - -
Dam A sandstone
up and downstream shoulder.

clean ?, Spread and levelled by Zone 3 used as up and downstream


suspect coarse
Glenbawn - main dam brokedown tractors with rakes. shoulder fill. Indication of breakdown at top
3 limestone sized and gap - 1.8 poor - - - - - - - - -
(prior to raising) on Watered in early stages of each layer on trafficking. Drp placed and
graded.
trafficking but discontinued. poorly compacted in mid to upper regions.

MW to FR rock, of very high UCS (except


reasonable
3B = 1.0 4p 10t SDVR, compacted MW dacite - high UCS). Zone 3B is the
Googong 3B & 3C dacite & granite - clean - (3C) to good - - - - - - - - -
3C = 2.0 dry predominant rockfill zone, 3C only used in
(3B)
the outer upstream slope.

No information, suspect reasonably to well


slate, chert and
Kisenyama 3 - - - - - - compacted given procedures used in Japan - - - - - - - - -
sandstone
at the time. Rock described as 'hard'.

No information, suspect reasonably to well


Kurokawa 3 - - - - - - - compacted given procedures used in Japan - - - - - - - - -
at the time.

Max size 1.0m Zone 3 used in up and downstream


4p 9t SDVR, no water reasonable to
La Grande 2 (LG2) 3 granite ? 8 to 19% < 19mm clean - 0.9 to 1.8 shoulders. Not sure if rock type is granite - - - - - - - - -
added good
< 3% fines. (assumed from geology).

Zone 3A used as rockfill transition in


0.9 to 1.2
max 0.45 m (3A), reasonable to downstream shoulder. The coarser Zone materials from Spec. > 95% reasonable to Limited information. Various materials from
Maroon 3A, 3B Porphery - - (3A) 10t SDVR, water added 4 - - 0.3 rollers or general traffic
0.9 m (3B) good 3B is the main rockfill zone. Rock excavation SMDD good excavation used, CH/MH clays excluded.
up to 1.5 (3B)
described as 'very hard'.

large portion from


1A (part of SM/ML - sandy Zone 1A used up and downstream of the
150 to 900 mm Spread and rolled by Used for both up and downstream Ignimbrite, highly Grid rollers (?), placed dry
Matahina 3 Ignimbrite ('hard') clean - 1.0 Poor. central silts to silty sands, - - 0.15 m Good central clay core (Zone 1). Sandy silts to silty
size. Very few tractor, no water added. shoulders. weathered of OMC.
core zone) 34% fines. sands, max. size 40 mm, 34% fines.
fines.
Appendix F Page F18

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 14 of 30)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Settlement
Pore Pressures in Core (Average Rockfill and/or Filters Comment Internal Monitoring
(from end of Steady / Instrument
Comments Comment mm/year) Comment Equipment
embankment Fluctuating Type/No.
mm % of H
construction)
Consider as normal, vertical
First filling occurred within the 3 year Negligible to limited positive PWP in core during 1.69 (0.7 to strains increasing with
Frauenau no information - not known - IVM 1254 - - -
monitored period construction. 3.6%) increasing depth below the
crest.
0 to 1% in upper part of core 800 mm or 1.6% (H1 H1 - 0 to 2.4% in rockfill, 2 to 4% in
Aug 1981 to Aug 1982. Delay
Fukada 4.8 to 5.8 years Steady close to FSL - - IVM (H2) 755 1.45% increasing to 3.0% in lower upstream), 595 mm or 1.3% core near base. IHM in downstream rockfill.
between EOC and start of first filling.
part. Normal. (H3 downstream) H3 - 0 to 2% (all in rockfill).

Subject to very rapid daily During construction - high PWP developed in the X-arm blocked with soil during
Rapid filling in 5 weeks (Jan to Feb Moduli of Zone 3A rockfill ~ 110 MPa
Geehi 0 to 0.1 years. Fluctuating drawdowns, up to 30 to 40 m over core, but dissipated quickly. Negligible by the end - IVM - - construction, tube filled with - HSGs and IHM in rockfill.
1966). at EOC (from HSG 1 & 2).
several days. of construction. concrete.

Relatively large settlements of


First filling started prior to end of shoulders during construction.
Annual seasonal drawdown. 1965 Construction - pwp up to 70 to 100% in core, but Most of stored
construction, reached within 48 m of IVM - 9 in total Arching across core confirmed by
Gepatsch -0.3 to 1.92 years Fluctuating and 1966 drawdowns were 54 and 80 rapid dissipation (reduced to 30% within 3 to 12 water is from - - limited information limited information IHM, IN.
FSL at EOC. FF completed by Oct (5 in core) pressure cells. A number of the
m respectively. months). snowmelt.
1966. vertical tubes were lost due to
considerable bending.

Started in August 1986 and reached 'normal' behaviour. Strains


Glenbawn, Saddle FSL in Oct 1990. FSL is 11 m below Not a annual drawdown. Significant increasing with depth from 0 to
0 to 4.1 years Fluctuating - 625 IVM 508 1.44% - - -
Dam A crest level due to storage capacity for drawdown in 1994-98 of 19m. 1% in upper 15 m to 4 to 5%
flood mitigation purposes. near the base of the core.

Started in August 1958 and reached 'normal' behaviour. Strains


Not an annual drawdown. Significant
Glenbawn - main dam FSL in May 1962. FSL is 15 m below During construction - positive PWP in core, r u up increasing with depth from 0.0
0.9 to 4.7 years Fluctuating drawdowns in 1964-66 of 23m, and 625 IVM 2425 3.21% - - -
(prior to raising) crest level due to storage capacity for to 0.4 in Zone 1A and 0.2 in Zone 1. to 0.5% in upper to 10 to 15 m
1979-81 of 26m.
flood mitigation purposes. to 5 to 7% in lower 50 to 75 m.

Likely to under-estimate the


Started just prior to EOC in early Greater settlements in core than in HSGs in core, filters and
Googong -0.1 to 1.21 years steady Maintained within ~ 3 m of FSL. - - HSGs in core 1180 1.90% settlement due to the 15 to 16 -
March 1977. To FSL by Aug 1978. filters and rockfill. rockfill
m spacing between HSGs.

Subject to very rapid drawdowns


Kisenyama 0 to 0.5 Filled within 0.5 years from EOC Fluctuating - - - - - - - - -
(daily ?), typically 5m, but up to 36 m.

Steady rise in reservoir to FSL over Subject to rapid drawdowns (daily ?),
Kurokawa 0 to 1.75 Fluctuating - - - - - - - - -
period of FF. typically 3 m but up to 20 m.

Started late Nov 1978 and reached


La Grande 2 (LG2) 0.12 to 1.2 years - No information beyond first filling - - - - - - - - -
FSL in late Dec 1979.

Not an annual drawdown, but


High PWP in core during construction (ru = 0.3 to 490 (H = 36) Vertical strain increases with
Steady rise to FSL from June 1973 to reservoir level does fluctuate.
Maroon 0 to 1.23 Fluctuating 0.75). Slow dissipation of excess PWP over 20 to - IVM, 3 to 1120 (H = 1.4 to 2.2% increasing fill depth from 0% - - -
mid 1974 Drawdowns slow and gradual. Lowest
25 years. 51) near crest to 4 - 6% at depth.
elevation = 18.3 m below FSL.

Greater settlement of rockfill


Close to FSL. Two large drawdowns shoulders than the core. Foundation
Matahina 0.11 to 0.24 years Nov 1966 to Jan 1967. Steady to undertake repairs, 1988 (18m) and Negligible construction pore pressures. - - - - - - in valley settled approx 180 mm HSGs and foundation plates.
1996-98 (18 m). during construction. Differential
settlement problems at abutments.
Appendix F Page F19

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 15 of 30)
SURFACE MONITORING - POST CONSTRUCTION
Crest Upstream Slope Downstream Slope
Time of Initial
Name Readings Comments
SLT (% per Comments Comments on Deformation and Observations.
(years after Settlement Displacement Settlement Displace- Settlement Displace-
log cycle of
EOC) (mm) (mm) (mm) ment (mm) (mm) ment (mm)
time)

Limited information. Displacement of the crest is downstream on first


150 mm or Consider as normal in terms of magnitudes. IVM profile post
Frauenau close to 0 (?) - 250 mm (3 yrs) - - - - filling, and accelerated in last 10 m of first filling. Crest spreading on Limited information. The core incorporates a central diaphragm wall.
0.18% (3 years) construction indicates uniform strain profile in core.
first filling of 60 mm.

130 mm or Displaced downstream on first filling (4.8 to 5.8 years). Low long- Limited data, but consider as 'normal'. Similar magnitude Consider as 'normal'. Very flat slopes for embankment height of 52 m. Rockfill is
Fukada 0 0.16 38 mm (7 years) - - - -
0.25% (7 years) term settlement rate. settlements between the core and weathered rockfill shoulders. weathered sandstone and granite.

236 mm or Consider as "normal" behaviour. Acceleration in rate of upstream


88 mm u/str edge Settlement of upstream crest and shoulder on FF of greater
0.27% u/str crest and slope settlement in 1989 to 1991 (23 to 25 yrs) is when
(33 yrs). 163 mm or 32 mm 154 mm or 153 mm magnitude than the downstream crest. Low long-term settlement Consider as "normal" behaviour. Reservior Level subject to numerous large and
Geehi 0.07 years edge. 142 mm 0.12 reservoir at low level and rising. Greater settlements on mid to
118 mm d/str edge 0.21% (33 yrs) (33 yrs) 0.25% (33 yrs) (33 yrs) rate. Displacement is downstream for all SMPs (upstream slope to rapid drawdowns.
or 0.16%, d/str lower d/str slope possibly due to greater thickness of dumped
(33 yrs) downstream slope), fluctuating with reservoir level.
edge (33 yrs) rockfill below SMP.

upstream moved Crest spreading of 600 to 750 mm at 2 yrs after EOC (upstream
Cracking in crest occurred as a result of the crest spreading and
upstream (500 mm, -665 mm crest and slope moved upstream and downstream crest and slope Settlement during construction indicates broad filters of significantly higher
0 years (EOC on 6 1650 mm or 732 mm or 1.0% 1087 mm or 886 mm differential settlement. Collapse settlement of the dry placed and
Gepatsch 0.83 13 yrs) and d/str (upstream) (22 moved downstream). Acceleration in deformation of upstream slope moduli than the rockfill and core, resulting in arching of core and high stresses in
Nov 1964). 1.08% (13 yrs) (22 yrs) 1.41% (22 yrs) (22 yrs) reasonably compacted rockfill on wetting contributed to the large
moved d/str (~ 200 yrs) on first filling and first drawdown. Foundation under the shoulders filter zones.
magnitude deformations of the upstream shoulder on first filling.
mm) contributed to the observed settlements.

On FF the crest and downstream slope displaced downstream, and Displacement of the upstream shoulder is potentially "abnormal",
0 (crest sett) the upstream slope displaced upstream. Post FF, the rate of the displacement rate to upstream continues at a steady to Consider deformation behaviour of the upstream shoulder as potentially
Glenbawn, Saddle 225 mm or 171 mm or -172 mm 66 mm or 0.31% 46 mm
0.47 (slopes and 0.16 -13 mm (12 yrs) settlement and displacement (to upstream) of the upstream shoulder increasing rate (in log time). The deformation behaviour of the 'abnormal'. The deformation behaviour of the crest, downstream shoulder and
Dam A 0.70% (12 yrs) 0.80% (12 yrs) (12 yrs) (12 yrs) (27 yrs)
crest displ) is greater than the crest and downstream shoulder. Other SMPs on crest and downstream are considered "normal". Embankment in internal core are considered "normal".
the upstream shoulder show similar trends. good condition (LWC 1995).

upstream edge The magnitude of crest spreading on FF is large. IVMs A and B


Large settlements and displacements on first filling. The upstream
moved upstream indicate large vertical strains at close to the interface between the
Glenbawn - main dam 760 mm or 695 mm or -260 mm 450 mm or 260 mm crest and slope displaced upstream and the downstream crest and Embankment raised by almost 24 m to 100 m in 1985-87. Only the deformation
0 years 0.26 (315 mm) and d/str core and filters (both up and downstream), due to differential
(prior to raising) 0.99% (27 yrs) 1.09% (27 yrs) (27 yrs) 0.71% (27 yrs) (27 yrs) slope displaced downstream. Records indicate crest spreading of of the original embankment has been considered here.
moved d/str (330 settlement as a result of collapse settlement of the dry placed and
more than 500 mm.
mm), 27 yrs poorly compacted rockfill.

1993 inspection descibed the embankment as having good Embankment raised by 4.5 m in 1991/92 to accomodate PMF analysis. Only the
0 (settlement) 233 mm or Limited records. Crest displacement is downstream on first filling and
Googong 0.20 79 mm (13 yrs) - - - - alignment with no visible abnormalities. Consider deformation deformation behaviour prior to raising has been analysed. The embankment was
0.94 (displ) 0.37% (13 yrs) remained virtually steady from 1980 to 1990.
behaviour as 'normal', but have limited records. also over-topped during construction.

Large downstream displacement of the crest on FF (180 mm), most


no information, 369 mm or Acceleration in downstream displacement at 1 year was possibly Constructed in narrow river gully with 26 to 38 degree abutment slopes.
87 mm of which occurred when the reservoir was raised above 80% of the
Kisenyama suspect close to 0.42% (16.5 0.29 421 mm (16.5 yrs) - - - rainfall related. Displacement pattern follows reservoir level with Consider deformation behaviour as "normal", although downstream displacement
(3 yrs) embankment height. Acceleration in crest displacement (to
EOC years) general trend toward downstream. is relatively large.
downstream) at 1 year, but not related to reservoir fluctuation.

no information,
323 mm or
Kurokawa suspect close to 0.17 258 mm (13 yrs) - - - - Crest settlement and displacement profiles show normal behaviour. Crest deformation considered normal Constructed in broad river gully with steep (33 to 47 degree) abutment slopes.
0.33% (13 years)
EOC

Settlement of the upstream crest and both the up and downstream Significant longitudinal cracking occurred during the latter stages
840 mm or
shoulders was relatively large on first filling, thereafter the long-term of FF. Cracking located on upstream side of the crest and at the Consider as a case of 'abnormal' deformation during and shortly after the period
0.56% (5.7 yrs)
0.13 years (sett) rate of settlement was normal cf similar materials. Large differential centre (mid 1979, length = 350m, width = 50mm). Late Sept 79 - of FF. Monitoring data indicates likely shear development in core that developed
La Grande 2 (LG2) 500 mm 0.135 675 mm (5.7 yrs) - - - -
0.22 yrs (displ) settlement at crest at the main section of 500mm. Large 300mm differential settlement across crack in central crest region during first filling, which is probably associated with collapse compression on
differential at
downstream crest displacement on FF, limited displacement (width = 150mm). Sharp tilt in inclinometer on u/str edge of crest wetting of dry placed rockfill.
main section.
thereafter. at 20m depth indicative of shear.

Settlement shows normal trend. Displacements are small (< 25 mm)


0 for all except and all displace to upstream. Long-term the rate of the displacement Weighting berms added during construction to upstream and downstream
269 mm or 94 mm or 0.28% -22 mm 114 mm or -18 mm
Maroon crest displacement 0.46 -15 mm (20 yrs) trend is virtually zero. The post construction internal settlement Deformation behaviour considered 'normal'. shoulders due to pre-sheared surfaces encountered in foundation along basalt
0.56% (20 years) (20 years) (20 yrs) 0.34% (20 years) (20 yrs)
(0.44 yrs) profile shows the settlement is distributed relatively uniformly over sills.
the full depth of the core.

Deformation records used for Matahina dam were up to the time of the
206 mm or 243 mm or Deformation records used were those prior to the earthquake in
484 mm (20 yrs, to earthquake in 1987. During the earthquake the downstream shoulder settled
0.26% (20 yrs, to 0.52% (20 yrs, to 1987. Large downstream displacement of the crest on first filling
Matahina 0.1 years 0.11 1987 prior to - - - Limited post construction data records. some 100 mm, the core (at crest level) settled approx. 50 to 100 mm and the
1987 prior to 1987 prior to (212 mm). Greater post construction settlements of rockfill shoulders
earthquake) upper upstream shoulder settled 800 mm (Penman 1988). Matahina also had
earthquake) earthquake) than core. Limited data records.
problems with piping adjacent to the abutments.
Appendix F Page F20

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 16 of 30)
GENERAL DETAILS CONSTRUCTION / DESIGN
Embankment
Construction Timing Dimensions
Classification
Name Owner/ References
Location Geology Foundation Dam Core Upstream Downstream Comments on Design
Authority Year Time Core Height, Length, Ratio
Zoning Width / Slope Slope
Completed (years) n Type H (m) L (m) L/H
Class Slope (H to V) (H to V)

Igneous and Volcanic - Medium sized central core with thin filter zones and dumped
US Corp of Cary (1958)
Mud Mountain USA, Washington andesite, volcanic Rock. 1941, Dec 0.75 5,1,0 c-tm SM/GM 128 213 1.7 2H to 1V 1.9H to 1V rockfill shoulders. Outer slopes ranged from 1.75 to 2.25 H
Engineers Cooke & Strassburger (1988)
breccia, tuff to 1V. Constructed in narrow, steep sided gully.

Water
Narrow central core supported by compacted rockfill Karasawa et al (1994)
Resources Granite with liparite and 3
Naramata Japan Rock (?) 1987 5,1,0 c-tn GM/GC 158 520 3.3 2.7H to 1V 2.05H to 1V shoulders. Single filter zones up and downstream of core. Okubo et al (1988)
Development dolerite intrusions. (seasonal)
Relatively flat upstream slope. Oikawa et al (1997)
Public Corp.

Sedimentary - cemented Thin central core supported bypoorly compacted rockfill Gamboa & Benassini (1967)
Core and downstream shoulder on
conglomerates and shoulders. Upper slopes at 1.75 H to 1V. Dam axis arched Marsal & Alberro (1976)
Netzahualcoyotl Mexico - bedrock. Upstream shoulder on dense 1964, Nov 1.5 5,2,0 c-tn MH 132 480 3.6 2H to 1V 2H to 1V
dense sandstones, to upstream. Built in a valley with narrow river section and Casales (1976)
river alluvium.
folded moderately steep abutment slopes. Wilson (1973)

Central core supported by poorly compacted rockfill


Goulburn shoulders. Downstream slope is a series of benches at Jacobsen (1965)
Nillahcootie Australia, Victoria granite Rock 1967, May - 5,2,0 c-tm CL 35 791 22.6 1.5H to 1V 1.55H to 1V
Murray Water 1.33H to 1V. Upper 9 m of upstream slope at 1.33H to 1V. SMEC (1996a)
Cofferdam at upstream toe.

Broad central clay core (1H to 1V upstream, 1.2H to 1V


Rock - upstream half of core and
Tennessee downstream) with thin filters and dumped and sluiced rockfill Leonard and Raine (1958)
Nottely USA, Tennessee - rockfill zones. In-situ soils left in place 1942, Jan 0.8 to 1.0 5,1,1 c-tk CL 56 592 10.6 1.75H to 1V 1.75H to 1V
Valley Authority shoulders. Outer slopes in series of benches sloped at 1.4H Blee and Meyer (1955)
under the downstream section of core.
to 1V.

Medium width central core supported by shoulders of


River section and left abutment - core
Metamorphic - quartzite 1.33H to 1V compacted rockfill. Weathered rockfill used in downstream Mitchell et al (1968)
Hydro Electric and shoulders mainly on gravels (talus
Parangana Australia, Tasmania and schist. Glaciated 1968, June 1.5 5,2,2 c-tm SM 53 189 3.6 1.33H to 1V (upper 20m), shoulder underlain by a filter. Central cut-off (5 to 15m Paterson (1971)
Commission & glacial drift) with cut-off to periglacial
valley. then 2.5H to 1V depth) to periglacial deposits. Cofferdam incorporated into Gerke and Hancock (1995)
deposits. Right abutment on bedrock.
the upstream toe.

Australia, QDNR (now Volcanic (andesitic) and QDPI (1994)


Peter Faust Rock 1990, Oct - 5,2,0 c-tn CL 51 530 10.4 1.6H to 1V 1.6H to 1V Thin central clay core with well compacted rockfill shoulders.
Queensland Sun Water) Igneous (Dioritic) QDNR (1998)

Portland Basalt interbedded with Core, transitions and downstream


Thin core aligned very slightly to upstream. Dam axis has
Round Butte USA, Oregon General Electric stratified siltsones and rockfill on bedrock. Upstream rockfill 1964, July 2 5,2,0 c-tn (u) SM 134 440 3.3 1.8H to 1V 1.7H to 1V Patrick (1967)
upstream curvature. Outer slopes steeper in upper 15 m.
Co. sandstones on shallow river bed deposits.

Variable, from bedrock to 40m of fluvio-


Metamorphic - quartzites Thin central core aligned slightly to upstream (slopes of Mitchell et al (1968)
Hydro Electric glacial deposits, due to ridge of bedrock
Rowallan Australia, Tasmania and schists. Glaciated 1967, Jan 2 5,1,0 c-tn GM 43 579 13.5 1.33H to 1V 1.45H to 1V 0.44H to 1V upstream and vertical downstream) supported Mitchell & Fitzpatrick (1979)
Commission in floor and abutments. Mostly on fluvio-
valley. by shoulders of compacted rockfill. Forza et al (1993)
glacial gravels.
Narrow central core (0.15H to 1V up and downstream) with
Sakamoto et al (1994)
Sagae Japan Diorite and Granite Rock 1990 unk 5,2,0 c-tn - 112 510 4.6 2.6H to 1V 2.3H to1V dual filters up and downstream. Relatively flat upstream and
Nose (1982)
downstream slopes.

Narrow central core aligned slightly to upstream with single


Seto Japan - - Rock 1977 unk 5,1,0 c-tn - 102 343 3.4 2.5H to 1V 2H to 1V filter zones up and downstream of the core. Relatively flat Kisa & Fukuroi (1994)
upstream slope.

Kansai Electric Narrow central core with single filter up and downstream. Kisa & Fukuroi (1994)
Shimokotori Japan - Rock 1972 unk 5,1,0 c-tn - 119 289 2.4 2.4H to 1V 1.85H to 1V
Power Co. Relatively flat upstream slope. Matsui (1976)
Appendix F Page F21

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 17 of 30)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Relative
Layer
ASCS Ratio or Dry Moisture ASCS Density or Level of
Zone Source n Placement Methods Comment Zone Source n Thickness Plant Comment
Class Density Content Class Dry Density Compaction
3 3 (mm)
(t/m ) (t/m )
Gravelly (?) -
Well compacted - 150 mm Silty sandy gravels, max. size 150 mm, compacted by caterpillar reasonable to Thin filter zone used up and downstream of the
Mud Mountain 1 - SM/GM - - 2 (filter) diorite, crushed max. size 100 - 300
layers, sheepsfoot rollers 20% fines. tractor good central core.
mm.

Well compacted filter zone of high moduli, used


Well compacted - 300 mm
101% SMDD 0.8% dry of OMC Well compacted core of processed weathered rhyolite 3 up and downstream of the central core. Width
Naramata 1 breccia GM/GC (?) 3 layers, 8p 18t vibrating 2A (filter) - 2.12 t/m 300 4p 16t vibrating rollers Good
2.02 t/m (9.6%) rockfill. and HW granite approx. 10 - 13 m upstream and 12 increasing to
tamping rollers
25m downstream.

Well-compacted filters. Fines content < 2 to 3%.


residual soil MH - sandy 92.5 - 94.5% Reasonably to well High plasticity sandy silts/clays placed Dual filter used downstream of the core and
2A & 2B (filter river alluvium (2A), SP (2A) 250 (2A) to 500 4p 3.5t SDVR, wetted. 2B
Netzahualcoyotl 1 (weathered silts/clays of SMDD compacted - 200 mm (loose) 7 to 8% wet of OMC well wet of OMC. Of stiff strength - Good single filter (Zone 2B) used upstream of the core.
3 / transition) & rockfill (2B) GW/GP (2B) (2B) also used as haulage path.
conglomerate) high plasticity 1.6 - 1.7 t/m layers, 6p 3.5t SDVR consistency (Su ~ 50 kPa). Width increasing with depth, up to 10 to 14 m at
the base.

Silty and sandy clays. Suspect 2A, 2B, 3B washed sand & Suspect reasonably to well compacted. Single
CL - silty and 100% SMDD Reasonably to well 450 (2A)
Nillahcootie 1 alluvial / colluvial 3 ~0.3% dry OMC reasonably to well compacted (practice at (filters / gravel (2A) and GP/GW (2A) - compacted - filters zones up and downstream of the core.
sandy clays 1.85 t/m compacted (?) 900 (2B)
the time). transition) crushed rock (2B) Zone 3B is a "compacted" rockfill transition zone.

limited information. Dual filters used upstream


Well compacted (?) - 150 to Sandy clays. Foundation filter used
placed at close to 2A & 2B and single filter used downstream of core. Filter
Nottely 1 - CL - sandy clays - 200 mm layers, sheepsfoot under downstream 1/3 of core along the quarried rock ? - - - - -
OMC (filters) extended under the downstream third of the
rollers full embankment length.
core.

1B - decomposed 1B - SM, silty Zone 1B used as main core material - Single filter used upstream of central core (2.5 to
Well compacted - 225 mm Spec. 1B - 1% dry to 10t SDVR
granodiorite sand (24% fines) Spec. > 98% silty sand with low plasticity fines. Zone 2A & 2B crushed dolerite 2A - GP reasonable to 6m wide) and dual filter used downstream of
Parangana 1B & 1A (loose) layers, 36t rubber 1% wet of OMC; 1A - - 450 to 900 450mm layers - 2p
1A - doleritic clay 1A - CL, sandy SMDD 1A used in cut-off trench - medium to (filters) river gravels 2B - GW good core (7 to 10m width) and on the foundation of
tyred rollers 1.5% dry to 1.5% wet. 900mm layers - 4p
(debris) gravelly clay high plasticity clay. the downstream shoulder.

alluvial sands & 1p 10t SDVR (2A & 2B)


Spec > 97% Well compacted - 150 mm Spec. 1% dry to 1% 2A, 2B & 2C SP/SW (2A&2B) Dual filters / transitions up and downstream of
Peter Faust 1 alluvium CL - sandy clays Gravelly sandy clays of medium plasticity. gravels, & crushed D R = 60 to 70% 375 2p 10t SDVR (2C) Good
SMDD layers, pad foot rollers wet of OMC (filters) GP (2C) the core.
rock (2C) moisture conditioned

Triple filter / transition zone used up and


SM -silty gravelly Well compacted - 305 mm Silty gravelly sands / sandy gravels. 2A, 2B, 2C processed sands 9t SDVR, no moisture
Round Butte 1 alluvium - - - - 305 Good downstream of the core, of moderately thin
sands layers, 45t pneumatic roller Average 30% non-plastic fines. (filters) and crushed rock conditioning
width.

Silty sandy gravels (23% fines, 62%


Spec. > 98% 2 - fine rockfill 2A - GW (< 5% Single filter used upstream (Zone 2) and
GM - silty sandy Well compacted - 450 mm Spec. OMC to 3% gravel) with low plasticity fines. Contact 3 2A - 450 reasonable (2) and
Rowallan 1 glacial till (doloritic) SMDD 2 & 2A (filters) 2A - fluvio-glacial fines, max. size 2A - 2.0 t/m 2p 10t SDVR downstream (Zone 2A) of central core.
gravel 3 layers, 8-12p 10t SDVR wet of OMC zone with foundation is a high plasticity 2 - 1350 (?) good (2A)
2.32 t/m gravels 300mm) Downstream filter of 2m minimum width.
sandy clay.
No information. Suspect well compacted
No information. Suspect well compacted. Dual
Sagae 1 - - - no information - given typical practices used in Japan at 2 (filter) - - - - - -
filters used up and downstream of the core.
this time.

No information. Suspect well compacted No information. Suspect well compacted given


Seto 1 - unk - no information - given typical practices used in Japan at 2 (filter) - - - - - - typical practices used in Japan at this time.
this time. Filters up and downstream of the core.

No information. Suspect well compacted


No information. Suspect well compacted given
Shimokotori 1 - unk - no information - given typical practices used in Japan at 2 (filter) - - - - - -
typical practices used in Japan at this time.
this time.
Appendix F Page F22

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 18 of 30)
MATERIALS
Rockfill Zone Rockfill Zone (in some cases earthfill stabilising berm)
Name
Dry Layer Level of Dry Layer Level of
Grading Clean / Grading Clean /
Zone Source Density Thickness Placement Method Comp- Comment Zone Source Density Thickness Placement Method Comp- Comments
Description Dirty 3 Description Dirty 3
(t/m ) (m) action (t/m ) (m) action

dumped and sluiced, 100


Coarse, max 1.2m
Mud Mountain 3 Andesite and tuff clean ? - 12 to 150% by volume water poor Very high strength rock. - - - - - - - - -
45 - 55% < 550 mm
to rock

Well-compacted rockfill. Karasawa et al


rhyolite and 3
Naramata 3A & 3B - - 2.05 t/m 1.0 4p 16t SDVR Good (1994) indicate moduli is high for rockfill, - - - - - - - - -
granite
190 MPa.

Spread by D8 and
max = 100 mm
conglomerate, trafficked. Water addition poor to
Netzahualcoyotl 3 55% < 19 mm dirty - 0.5 to 1.0 Low strength conglomerate used as rockfill. - - - - - - - - -
low strength specified and heavy rains reasonable.
4% fines
during construction.

metamorphosed dumped ?, no indication


No information on placement methods,
Nillahcootie 3A sandstone and - - - - of lift thickness or if water poor (?) - - - - - - - - -
suspect placed without formal compaction.
mudstone added during placement.

quartzite with Rock described as "sound, durable rock of


Nottely 3 - - - 9.1 dumped and sluiced poor - - - - - - - - -
20% mica schist 'good' compressive strength".

high degree of
4p 10t SDVR, no water Quartzite described as 'extremely hard'. schist (HW to variability 4p 10t SDVR, no water
max size ~0.9m reasonable to reasonable to Zone 3B used in downstream shoulder.
Parangana 3A quartzite, sound clean ? - 1.35 addition required in Zone 3A used in upstream shoulder as well 3B MW), and gravelly max ~ 900 mm dirty - 0.9 addition required in
<20% < 12.5mm good good Lesser quality rockfill than Zone 3A.
specification. as on the outer downstream shoulder. talus 0 to 90% < 19mm specification.
< 5% fines.

hornfelsed 1.5
6p 10t SDVR, 20% by reasonable to
Peter Faust 3 andesite & - clean ? - (0.75 within 3 Rock of very high UCS. Well compacted. - - - - - - - - -
volume water added. good
andesitic tuff m of filter)

4p 9t SDVR, no water Suspect relatively small sized rockfill given


Round Butte 3 basalt - - - 0.61 good - - - - - - - - -
added placed in 0.6 m lifts. Dry placed.

max size 915mm Zone 3A used in up and downstream


4p 10t SDVR, no water
<25% pass 20mm 1.93 to 2.08 reasonable to shoulder. Grading, and relative density
Rowallan 3A quartzite, sound clean 3 1.35 addition required in - - - - - - - - -
very low fines t/m good variation between lower and upper halves of
specification.
content layer.
No information. Suspect well compacted
Sagae 3 - - - - - - - given typical practices used in Japan at this - - - - - - - - -
time.

No information. Suspect well compacted


Seto 3 - - - - - - - given typical practices used in Japan at this - - - - - - - - -
time.

No information. Suspect well compacted


Shimokotori 3 - - - - - - - given typical practices used in Japan at this - - - - - - - - -
time.
Appendix F Page F23

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 19 of 30)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Rainfall Other Types of
Name Timing Settlement
Pore Pressures in Core (Average Rockfill and/or Filters Comment Internal Monitoring
(from end of Steady / Instrument
Comments Comment mm/year) Comment Equipment
embankment Fluctuating Type/No.
mm % of H
construction)

By 1958 had not yet reached FSL, Annual drawdown, 20 to 35 m in first precipitation
Mud Mountain 0 to >17 years Fluctuating - - - - - - - -
still 30 m below FSL. two drawdowns. mostly as snow.

Small strains for dam of this


Started Oct 1988 and reached FSL in
1 to 1.75 years (not sure height, 0 to 0.5% near crest Likely high moduli of dry, well
Naramata May 1989, but not sure of timing with - not known - heavy snow falls IVM 2920 1.85% - -
of timing) increasing to 2 to 2.5% near compacted core.
respect to end of construction.
base.

Started late May 1966. Within 15 m normal, vertical strain


IVM in rockfill. Vertical
of FSL in Jan 1967. Upstream During construction - high PWP in core. ru approx increasing with depth. Up to
strains in rockfill greater
Netzahualcoyotl 1.5 to > 4.8 years rockfill flooded during construction - - 0.45 based on estimated superimposed load, but 2500 IVM 2970 2.72% 1.5% in upper 20 m increasing - -
magnitude than in the core
when reservoir level rose above the will be higher when arching effects are considered. to 4% at 90 m deth below
at similar depths.
crest of the cofferdam in Sept 63. crest.

Annual drawdown, typically between


0 to 1.3 years, possibly At FSL in August 1968 (1.3 years),
Nillahcootie Fluctuating 4 and 8 m. Larger drawdowns in - 985 - - - - - - -
less. which was start of records.
1983 (10 m), and in 1991 (11.5 m).

Jan 1942 to July 1942 (vitually at Annual drawdown. Lowest level is


Nottely 0 to 0.6 years Fluctuating - - - - - - - No monitoring during construction -
FSL). virtually empty.

351mm 0.96% to 1.41% Vertical strains of 0 to 1% in


IVM, 4 (all
Generally remains close to FSL. One (H=36.5m) to (greater strains the upper 10 to 20m increasing
Parangana 0.6 to 0.74 years First filling from Jan to Feb 1969 Steady - 1250 close to - - -
drawdown event in late 1969 of 26m. 733mm for increasing to 2.5 to 3.5% at about 40 to
centreline)
(H=52m) height) 50m depth. Some variability.

Stored water during construction to During construction - high PWP in the contact
Fluctuating Slow reduction in reservoir level since
Peter Faust -0.85 to 0.42 years within 22 m of FSL. Filled very zone, otherwise limited to negligible positive PWP - - - - - - - -
(slow) reached FSL (10 m in 6 years).
quickly in mid to late Dec 1990. (ru approx 0 to 0.1).

Filling started before end of


-0.6 to 0.6 (Jan 1964 to Relatively steady once first filling Monitoring of SMPs on slopes
Round Butte construction. Within 20 m of FSL at Steady - - - - - - - -
Jan 1965) completed started before EOC.
EOC.

Not an annual drawdown, but


drawdowns are relatively rapid and
Rowallan 0.38 to 0.7 years First filling from April to Oct 1967 Fluctuating - 1780 - - - - - - -
typically range from 8 to 30m, largest
in 1979 of 36 to 38m.

4000 (mostly as Increasing strain with depth.


Sagae - not known - not known - IVM 2430 2.2 (1.0 to 5.4%) - - -
snowfall) Considered normal.

Seto 0 to 0.45 years - Fluctuating rapid daily drawdowns up to 34 m. - - - - - - - - -

Shimokotori 0 to 1.25 years - Fluctuating large annual drawdown of 30 to 40 m. - - IVM 3850 3.25 (2 to 4.8%) - - - -
Appendix F Page F24

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 20 of 30)
CONSTRUCTION
SURFACE MONITORING - POST CONSTRUCTION
Crest Upstream Slope Downstream Slope
Time of Initial
Name Readings Comments
S LT (% per Comments Comments on Deformation and Observations.
(years after Settlement Displacement Settlement Displace- Settlement Displace-
log cycle of
EOC) (mm) (mm) (mm) ment (mm) (mm) ment (mm)
time)

Differential settlement between core and rockfill resulted in


Significantly greater magnitude settlement of the rockfill shoulders Main embankment constructed from March to December 1941. Outlet works not
823 mm or cracking along junction of the core and transition on both upstream
Mud Mountain 0 - 274 mm (8.5 yrs) - - - - (both up and downstream) than the crest above the core, approx. 450 completed until 1948 due to war interruption. Likely collapse settlement of
0.66% (8.5 yrs) and downstream sides. No measurable differential settlement
mm greater. dumped and sluiced rockfill shoulders.
since 1948.

177 mm (on first Only information is 177mm downstream displacement of crest on


Naramata not known - - - - - - Only have deformation during construction. IVM indicates vertical deformation of
filling) first filling. Most of displacement occurs when reservoir level above -
core is normal.
85% of FSL.

Large magnitude settlement of the crest and upstream slope prior to


596 mm or and on FF. The IVM at the crest indicates a large vertical differential Consider as 'normal' deformation behaviour. IVMs indicate shear
0.08 (crest) 0.45% (5.6 yrs), at the upstream core/filter interface of 300 mm (greater shoulder type deformation at the upstream interface between the filter and
471 mm or 303mm or 73 mm Consider as case of 'normal' deformation. IVM data indicate shear deformation
Netzahualcoyotl 0.2 (d/stream) I3 (u/str edge) = 0.23 - - settlement), and is reflected in the SMP data. The crest initially core. No observation of crest cracking reported. It was likely that
0.41% (5.6 yrs) 0.24% (5.6 yrs) (2.2 years) at the upstream core / filter interface during first filling.
0.36 (upstream) 760mm or displaced upstream in the early to mid stages of FF then to some collapse settlement of the upstream rockfill occurred during
0.70% (2.75 yrs) downstream in the later stages when within 15m of FSL (similar FF.
pattern for the upper downstream shoulder).

1.3 year delay to monitoring, which missed the period of FF.


Likely that rockfill was susceptible to collapse compression on
Displacements indicate post FF crest spreading (~ 120 mm), with
66 mm or 0.19% 171 mm or -42 mm 103 mm or 104 mm wetting, which is likely to have affect the lateral support afforded to
Nillahcootie 1.3 years. 0.21 34 mm (16 years) upstream slope moving upstream. Acceleration in settlement and Potentially 'abnormal' behaviour on drawdowns of more than 9 m.
(16 years) 0.50% (31 years) (31 years) 0.39% (31 years) (31 years) the central core. Potentially 'abnormal' behaviour on drawdowns
non-recoverable upstream displacement of the upper upstream
of more than 9 m.
shoulder on large drawdown.

Up to 1955, "severe" longitudinal cracking on crest coincident with


SMPs in centre of crest. Up to 1955 (13 years after EOC) the the upstream edge of core. In 1955 longitudinal cracking more
310 mm or upstream edge of the crest had settlement more than 250 mm more toward the centre of the crest. Initial cracking probably attributable
0.33 (4 month after Suspect deformation behaviour of upstream crest and shoulder is potentially
Nottely 0.56% (10.3 0.47 62 mm (3 years) - - - - than the centre of the crest. Increase in settlement rate (rate in to collapse settlement on wetting of the dumped and sluiced
EOC) "abnormal". The embankment is subject to very large drawdowns.
years) terms of log time) from 5 years after EOC. Crest displaced upstream rockfill. The increased settlement rate may be related to
downstream on first filling. softening and cracking of the core due to poor lateral support from
the rockfill.

SMPs only on the crest of the embankment. Magnitude of


0 yrs (sett) 111mm or Surveillance review in 1995 commented that the embankment was
Parangana 0.05 32mm (26.5 yrs) - - - - settlement is small and the long-term rate low. Acceleration of Reservoir level is steady at close to FSL. The embankment has a relatively
0.5 yrs (displ) 0.21% (26.5 yrs) in a satisfactory condition. Deformation is considered as 'normal'.
settlement rate during large drawdown in late 1969. Crest steep mid to upper upstream slope.
On the large drawdown in 1969 the crest settled 15 to 20 mm.
displacement is downstream on FF (very low rate long-term).

0.34 years (close Small post construction deformations. Deformation monitoring Consider deformation behaviour as 'normal'. Delay of 0.34 years
51 mm or 0.10 18 mm or 0.05 2 mm Consider as 'normal' deformation. Post construction deformations of small
Peter Faust to FSL when 0.12 -9 mm (6.5 yrs) - - virtually missed first filling. Displacements less than 10 mm in 6.5 before the start of monitoring, during which time first filling virtually
% (6.5 yrs) % (6.5 yrs) (6.5 yrs) magnitude.
started) years. Low long-term settlement rate. completed.

Longitudinal cracking on the crest observed as reservoir level


0 for settlements Greater settlement of rockfill shoulders than core. Crest spreading approached FSL, gradually widened to 15 mm (filled by sluicing Collapse settlement of upstream rockfill on first filling due to dry placement (even
161 mm or 331 mm or 367 mm
Round Butte 0.32 for displ of 0.075 569 mm (1.3 yrs) - - occurred on FF (> 130 mm). Large downstream displacement of sand into crack). Crack located in area of transition on the though reasonably to well compacted) contributed to the differential settlement
0.12% (1.3 yrs) 0.33% (1.3 yrs) (1.3 yrs)
u/str crest crest during the latter stages of FF. downstream side of the core. Suspected differential settlement and cracking. Tailwater impounds the lower part of downstream slope.
and cracking at upstream core interface.
Data for SMPs on crest only. Settlements show small magnitude
107mm or Surveillance review in 1995 commented that the embankment was
and low long-term rate. Crest displacement at main section is Consider deformation behaviour as 'normal'. Piping incident occurred shortly
Rowallan 0.22 yrs 0.095 60mm (23 yrs) - - - - in a satisfactory condition. Deformation behaviour is considered
0.27% (23 yrs) downstream on FF, with continued downstream trend post FF at a after first filling in 1968.
as 'normal'.
gradually reducing rate.

Sagae - - - - - - - - Only have IVM data during construction. Deformation profile Dam used for flood control, municipal water supply and power generation. Only
No post construction deformation data.
considered normal. has internal settlement records for the core during construction.

no information, Relatively large settlement on and shortly after FF (150 mm).


359 mm or Limited information. Settlement and displacement behaviour Constructed in narrow river gully with 39 degree abutment slopes. Deformation
Seto suspect close to 0.22 201 mm (9 years) - - - - Downstream displacement on FF of 120 mm, occurred when
0.35% (9 years) considered "normal" . behaviour considered normal.
EOC reservoir raised above 80% of embankment height.

no information, Relatively large settlement on FF (300 mm), most of this occurred in Acceleration in crest settlement on first 3 drawdowns and on
542 mm or Constructed in narrow river gully with steep (40 to 47 degree) abutment slopes.
Shimokotori suspect close to 0.25 139 mm (13 years) - - - - the upper 40 m (from IVM). Downstream displacement on FF of several later drawdowns. Magnitude of settlement on drawdown of
0.46% (13 years) Possibly "abnormal" deformation post construction given continued acceleration
EOC 95mm occurred when reservoir raised above 80 to 85% of 10 to 50 mm. Possible plastic deformation of the core on large
in crest settlement on drawdown.
embankment height. drawdowns, possibly within the upper 40m.
Appendix F Page F25

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 21 of 30)
GENERAL DETAILS CONSTRUCTION / DESIGN
Embankment
Construction Timing Dimensions
Classification
Name Owner/ References
Location Geology Foundation Dam Core Upstream Downstream Comments on Design
Authority Year Time Core Height, Length, Ratio
Zoning Width / Slope Slope
Completed (years) n Type H (m) L (m) L/H
Class Slope (H to V) (H to V)

Central clay core supported by well-compacted rockfill


QWRC (1987, 19XXb)
Australia, QDNR (now Interbedded quartzites & shoulders. Large step in upstream slope at about mid-
Split Yard Creek Rock 1980, Feb 2 5,2,1 c-tm CL 76 1141 15.0 1.75H to 1V 1.8H to 1V Rogers & Pearce (1991)
Queensland Sun Water) metavolcanics height. Zone of weathered rockfill used in the lower
QDPI (1995, 1996)
downstream shoulder.

Tennessee Sedimentary - shale and 2 Broad central clay core (0.75H to 1V up and downstream) Leonard and Raine (1958)
South Holson USA, Tennessee Rock 1950, Oct 5,2,1 c-tk CL 87 487 5.6 1.9H to 1V 1.8H to 1V
Valley Authority sandstone (seasonal) with thin filters and dumped and sluiced rockfill shoulders. Blee and Meyer (1955)

Sedimentary - limestone,
Core and filters on sound rock. Narrow central core with broad transition zones (up to 28 m
Srinagarind Thailand - sandstones, quartzites, 1978, May 2.3 5,2,0 c-tn GC/SC 140 610 4.4 2H to 1V 1.9H to 1V Champa & Mahatharadol (1982)
Shoulders on at least weathered rock. wide) and compacted limestone rockfill shoulders.
shales

Very narrow central core, slightly upstream sloping with dual


3 filters up and downstream. Embankment has a slight arch to Kjaernsli et al (1982)
Svartevann Norway - Granitic gneiss Rock 1976, Oct 5,2,0 c-tn (u) GM 129 420 3.3 1.6H to 1V 1.43H to 1V
(seasonal) upstream. Downstream slope is series of benches at 1.35H Dibiagio et al (1982)
to 1V.

Narrow central core with single filter up and downstream. JNCOLD (1976)
Taisetsu Japan, Hokkaido - Metamorphic - clay slate Rock 1975, Oct 7.5 5,1,0 c-tn - 86.5 440 5.1 2.65H to 1V 2H to 1V
Relatively flat upstream slope. Sakamoto et al (1994)

Central clay core with slight upstream slope (0.9H to 1V up


Australia, New and -0.2H to 1V downstream) supported by compacted Howard et al (1978)
Talbingo SMHEA Volcanics - rhyolite lavas Rock 1970, Oct 2.1 5,2,0 c-tm (u) CL 162 701 4.3 2H to 1V 2H to 1V
South Wales rockfill shoulders. Moisture of core material adjusted due to SMHEA (1998c)
high pore pressures observed in early stages of construction.

Gneiss and JNCOLD (1976)


Electric Power Central narrow core supported by broad filter/transition zones
Japan, Honshu conglomerate with 1.5 Kawashima & Kanazawa (1982)
Tedorigawa Development Rock 1978, Nov 5,2,0 c-tn GM/GC 153 420 2.7 2.6H to 1V 1.85H to 1V and compacted rockfill shoulders. Cofferdam at upstream
Island numerous porphyritic (seasonal) Sakamoto et al (1994)
Corporation toe. Upstream slope is relatively flat.
dyke intrusions. Nakagawa et al (1985)

Sedimentary
Rock - of lesser quality under rockfill Medium width core aligned slightly to upstream. Dual filters
Terauchi Japan, Fukuada - (argillaceous) and 1977, Feb 4.25 5,2,0 c-tm SC/SM 83 420 5.1 2.7H to 1V 2.1H to 1V Kanabayashi et al (1979)
shoulders than core. up and downstream of the core. Relatively flat slopes.
metamorphic (schist)

Central core (slopes of 0.3H to 1V up and 0.2H to 1V SMEC (unpublished)


Melbourne Sedimentary - sandstone
Thomson Australia, Victoria Rock 1983, July 2.5 5,2,0 c-tm SC 166 590 3.6 1.65H to 1V 1.75H to 1V downstream) supported by compacted rockfill shoulders. Fill MMBW (1987b)
Water Corp. & siltstone.
berm at upstream toe of right abutment. Connell Wagner (1998c)

Sedimentary, some
metamorphism - Thin central core with narrow, single filter zone (6 m wide) up Shiraiwa & Takahashi (1985)
Tokachi Japan, Hokkaido - Rock ? 1983, August 4 5,1,0 c-tm GC 84.3 406 4.8 2.6H to 1V 2H to 1V
keratophyre, clay slate, and downstream. Relatively flat slopes. Sakamoto et al (1994)
sandstone

Narrow central core with dual filter / transition zones up and


Volcanics - tuff and downstream and rockfill shoulders. No indication of degree
Upper Dam Korea - Bedrock ? 1984, Sept 1.5 5,2,0 c-tn GC 88.5 269 3.0 2.1H to 1V 1.8H to 1V Hong et al (1994)
rhyolite ashy tuff of compaction of rockfill. Cofferdam incorporated into the
upstream toe of the embankment.
Appendix F Page F26

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 22 of 30)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Relative
Layer
ASCS Ratio or Dry Moisture ASCS Density or Level of
Zone Source n Placement Methods Comment Zone Source n Thickness Plant Comment
Class Density Content Class Dry Density Compaction
3 3 (mm)
(t/m ) (t/m )
3 Sandy clay to clayey sands of medium Limited information. Dual filters downstream of
CL, CL-SC - 1.6 to 1.82 t/m Well compacted (?) - 150 to
(Spec. 1% dry to 1% plasticity, well-compacted. 0.6 m thick 2A & 2B 2p 10t SDVR, watered prior core (width = 4.75 m), single filter upstream
Split Yard Creek 1 alluvial sandy clays to (Spec > 98% 200 mm layers, padfoot rollers alluvial 2A < 8% fines (Spec > 70% D R) 300 Good
wet of OMC) contact zone of wet placed medium to (filters) to rolling (Zone 2B). Zone 2B filter under the downstream
clayey sands SMDD) (Cat 815 & 825)
high plasticity clays. shoulder.

Well compacted (?) - 150 to Sandy or silty clays of medium plasticity.


possibly close to or 2A & 2B limited information. Dual filters of thin width (6.1
South Holson 1 - CL - 200 mm layers, sheepsfoot Upper 3 m compacted by trucks and quarried rock ? - - - - -
wet of OMC (filters) m wide) up and downstream of the core.
rollers tractor.

Mainly clayey sandy gravels, with 26 to 3 Well compacted filter / transition zones. Dual
GC/SC, but 100% SMDD Well compacted - 400 mm 2A & 2B (filter limestone (2A) & 2.1 - 2.2 t/m 3p SDVR (2A), 4p SDVR
Srinagarind 1 colluvial, lateritic 3 0.2% wet of OMC 42% medium plasticity fines (max. size GW 400 Good system used up and downstream of core and of
variable 1.85 t/m (loose) layers, 4p SDVR / transition) quartzite (2B) (Spec > 70% D R) (2B)
150 mm). moderate combined width. Fines content < 7%.

GM - silty sandy fluvial gravels (2A) 2p 8t SDVR (2A), 4p 13t Well compacted filters of moderate width (8 to 13
100.4% SMDD Well compacted - 500 mm Well compacted silty sandy gravel, max. 2A & 2B 500 (2A) and
Svartevann 1 morraine gravel, 20% 3 0.4% wet of OMC & crushed rock GW/GP - SDVR (2B). Wetted prior Good m wide) up and downstream of the core. Fines
2.15 t/m layers, 8p 8-10t SDVR size 170 mm. (filters) 1000 (2B)
fines (2B) to compaction. content < 5%.

No information. Suspect well compacted


No information. Suspect well compacted given
Taisetsu 1 - unk - no information - given typical practices used in Japan at 2 (filter) - unk - - - -
typical practices used in Japan at this time.
this time.

Overall average is at Zone 2A was of narrow width and placed only


CL (some CH, Gravelly clays to clayey gravels of
CW to HW OMC. Lower region downstream of core. Broad Zone 2B transition
SC & GC) - 101.6% SMDD Well compacted - 150 mm medium plasticity. Fines content 18 to 2A & 2B (filter 2A - SW/GW 4p 10t SDVR. Zone 2B
Talbingo 1 andesitic basalt, 3 wet of OMC, and crushed rock, fresh - 450 Good up and downstream of the core; narrow width at
mainly gravelly 1.84 t/m layers, sheepsfoot rollers 70%. Contact CL/CH clay placed slightly / transition) 2B - GW/GP was sluiced.
colluvium upper region slightly crest increasing to ~ 60 m (d/str) and 30 m
clays wet of OMC.
dry of OMC. (upstr) at the base.

Limited information, but suspect well compacted.


blended and Limited information, but typical practice is
GM/GC (max 3 Well compacted (?) - heavy 2A & 2B from crushed rock 3 Dual filters up and downstream of central core.
Tedorigawa 1 processed talus or 1.9 t/m Slightly wet of OMC for heavy rolling of materials processed or unk 2.04 t/m - - -
150 mm) rollers (filters) (?) Increasing width with depth, up to 60m wide at
rock blended from talus and weathered rock.
the base.

Well compacted - 200 mm GP/GW -


residual soil & SC/SM - gravelly Spec. OMC to 2% Well compacted gravelly silty sand core, 2A & 2B 3 limited information, suspect well compacted.
Terauchi 1 - layers, 12p Cat 825B pad foot - gravels and 2.15 t/m - - -
weathered schist silty sands wet of OMC 20 to 33% fines. (filters) Slightly sandy gravels and cobbles.
roller cobbles

Clayey sands with fines of medium Light compaction to reduce potential acrhing
SC - clayey Well compacted - 150 mm
99.9% SMDD plasticity. Fines content = 44% and clay 2A & 2B sandstone, 1p 10t SDVR, without effects in core. Estimated width of 6m
Thomson 1 residual granite sands, 44% avg 3 layers, 12p 19t sheepsfoot 2% wet of OMC - - 500 reasonable
1.79 t/m fraction = 15%. High pore pressures due (filters) indurated vibration downstream of core (2A and 2B combined) and
fines content rollers
to wet placement. 3m upstream (2A).

3
GC - clayey Well compacted - 200 mm Clayey sandy gravel core, max. size 50 2.13 t/m limited information. Max size = 200 mm. 6.5%
weathered clay > 95% SMDD Spec. 1% dry to 1% GP (slightly
Tokachi 1 sandy gravel 3 layers, 8p 33t compacting mm. Blended mixture to control fines 2 (filter) river alluvium (n = 21%, e = 400 - - moisture content at placement. Suspect well
slate and colluvium 2.07 t/m wet of OMC sandy gravel)
(15% fines) dozer (?) content. 0.27) compacted.

GC - clayey
2A - SC (fine
sandy gravels,
colluvium and 3 Limited information on materials and 2A & 2B sand and clay ?) 3
Upper Dam 1 max size 1.72 t/m no information 17.3% river alluvium 1.9 t/m - - - No information on placement methods of filters.
alluvium placement methods of the core. (filters) 2B - coarse
150mm (well
sand
graded)
Appendix F Page F27

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 23 of 30)
MATERIALS
Rockfill Zone Rockfill Zone (in some cases earthfill stabilising berm)
Name
Dry Layer Level of Dry Layer Level of
Grading Clean / Grading Clean /
Zone Source Density Thickness Placement Method Comp- Comment Zone Source Density Thickness Placement Method Comp- Comments
Description Dirty 3 Description Dirty 3
(t/m ) (m) action (t/m ) (m) action

Rock of high to very high UCS. Zone 3A


MW to FR rock. Zone 3C used in the lower
brecciated max 1.0 m spread by dozer, 5p 10t (fresh rock) used in upper region of the brecciated max 0.5 m dirty (broke spread by dozer, 5p 10t
1.84 to 2.25 downstream section. Susceptible to
Split Yard Creek 3A & 3B quartzites and <19 mm = <15% - - 1.5 to 3.0 SDVR, 5% by volume reasonable upstream shoulder, Zone 3B used in lower 3C quartzites and <50% pass 9.5mm down on 3 0.75 to 1.5 SDVR, 5% by volume reasonable
t/m breakdown on placement. Foundation filters
greenstone (3A) & <30% (3B) water added upstream shoulder and upper downstream greenstone <10% fines compaction) water added
used to control downstream seepage.
shoulder.

Sound, high to very high strength rock used


sandstone, some
South Holson 3 - - - 9.1 dumped and sluiced poor in downstream shoulder and poorer quality - - - - - - - - -
shale
used in upstream shoulder.

Max 0.7 m (3A), 1.5 3A - good


Limestone - hard 1.0 (3A), 2.0 Spread by D8, then 4p Zone 3B forms the outer 20% of the rockfill
Srinagarind 3A & 3B m (3B). clean ? (n = 22%) 3B - - - - - - - - - -
and durable (3B) 13.5t SDVR. zone.
5 to 25% < 19 mm reasonable

Rockfill placed without addition of water,


8p 13t SDVR, no water
Svartevann 3 granitic gneiss - - - 2.0 Reasonable only reasonably compacted. No indication - - - - - - - - -
added
of grading.

No information. Suspect well compacted


Taisetsu 3 - - - - - - - given typical practices used in Japan at this - - - - - - - - -
time.

rhyolite & Zone 3B of lesser quality than 3A and 3C,


rhyolite & Good quality rockfill. Zone 3A used as 3 good (3B),
3 porphyry 3C - clean? 3B - 2.07 t/m 3B - 0.9 m 4p 10t SDVR, Zone 3B possibly 'dirty'. Zone 3B used as outer
Talbingo 3A porphyry - clean 1.92 t/m 0.9 4p 10t SDVR, sluiced. Good inner rockfill zone on downstream side of 3B & 3C - reasonable
(3B - MW to FR, 3B - dirty ? 3C - 2.01 t/m 3 3C - 1.8 m was sluiced. rockfill in downstream shoulder. Zone 3C
(SW to FR) core. (3C)
3C - SW to FR) used in upstream shoulder.

3 Suspect well compacted given typical


Tedorigawa 3A & 3B - - - 1.98 t/m - SDVR - - - - - - - - - -
procedures used in Japan at this time.

max size 1.5m


lutaceous 3 Well compacted rockfill. Rock refered to as
Terauchi 3 2% fines clean 2.1 t/m 0.5 to 1.5 4p 22t SDVR good - - - - - - - - -
schist 'hard'.
Cu ~ 20

Sandstone rock of very high strength and


Sandstone and Sandstone and Zone 3C of lesser quality than Zone 3A
(bulk density = 4p 10.5t SDVR, no siltstone of high to very high strength. Zone 4p 10.5t SDVR, no reasonable to
Thomson 3A siltstone, SW to - clean ? 3 1.0 good 3C siltstone, MW to - dirty - 1.0 rockfill. Zone 3C used in outer downstream
2.13 t/m ) indication if water added. 3A used in upstream shoulder and most of indication if water added. good
FR FR shoulder.
downstream shoulder.

keratophyre (VH max 200 mm


3 Suspect well compacted given typical
Tokachi 3 strength), 30% 25% < 19 mm clean ? 2.1 t/m 1.0 - - - - - - - - - - -
practices used in Japan at this time.
slate Cu ~ 22

3 No information on placement methods of


Upper Dam 3A & 3B tuff and rhyolite - - 1.8 t/m - - - - - - - - - - - -
rockfill. Referred to as free draining.
Appendix F Page F28

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 24 of 30)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Settlement
Pore Pressures in Core (Average Rockfill and/or Filters Comment Internal Monitoring
(from end of Steady / Instrument
Comments Comment mm/year) Comment Equipment
embankment Fluctuating Type/No.
mm % of H
construction)

Normal vertical strain profile,


Rapid daily drawdowns of typically 5 PWP levels in core remain fairly constant. IVM - (ES1 - on
Not commenced until July 1983 (3.8 close to 0% in upper 20 m 3 INs to monitor horizontal
Split Yard Creek 3.8 to 5.7 years Fluctuating to 10 m. Largest drawdown (1995) of Insufficient time to respond to rapid daily - upstream side 1320 1.76 (1 to 6%) - -
yrs after EOC). increasing to 4 to 6% at 50 to movements
50 m. drawdowns. of core)
70 m depth.

Nov 1950 to mid 1952. By June annual drawdown of 6 to 27 m (1951


South Holson 0 to 1.62 years Fluctuating - - - - - - - No monitoring during construction -
1951 reached within 9 m of FSL to 1957).

First filling started prior to end of


construction (to within 70 m of FSL PWP in core during construction, maximum
Srinagarind -0.55 to > 2.9 years - - - - - - - - - -
before EOC). Not completed by April approx 45% of applied load.
1981

No EPWP in core during construction.


Late 1975 to end of 1978. Water
Post construction - PWP in core fluctuates with
Svartevann -0.9 to 2.23 years stored during construction, to within Fluctuating Annual drawdown of 15 to 20 m (?). - - - - - - - -
reservoir level but at lower and reducing amplitude
37 m of FSL at EOC.
across the core.

Taisetsu - not known - not known - - - - - - - - -

ES1 at main section (H =


154m), H = 98 m at ES2.
During construction - PWP in lower, wet placed 2410 (ES 2) HSGs indicate Zone 2B is of high
Started in May 1971 and reached 2.46% (ES 2) & Some variability but consider HSGs and IHMs in rockfill
Talbingo 0.6 to 1.3 years steady Typically within 5 m of FSL. region of the core were high (ru up to 0.81), slow - 2 (both in core) & - moduli compared with core and
FSL by Jan 1972. 3.40% (ES 1) normal. Low strains in lower zones.
dissipation with time. 5202 (ES 1) rockfill shoulders (3A to 3C).
region possibly due to arching
influence.

Large vertical strains in mid to


During construction - PWP up to 60 to 70% of
Started June 1979, reached FSL in lower region of core (up to 7 to
Tedorigawa 0.6 to 1.5 years - not known applied load. Relatively rapid dissipation during - HSGs in core 5860 3.96% - - -
March 1980 8%). Possible plastic
winter shutdown and after construction.
deformation of core.

Suspect fluctuating, only limited data


Terauchi 0 to 0.66 years completed Oct 1977 Fluctuating (?) post first filling showing one 15 m - - - - - - - - -
drawdown cycle.

During construction - reasonably high PWP during


IVMs installed in core but
Steady at FSL from 1991 to 1997 (7 construction, r u values up to about 0.5. Dissipated HSGs indicate limited diffferential HSGs, IHMs and INs -
Started July 1983. Reached FSL at became blocked during
Thomson 0 to 7.3 years Steady to 14 yrs), then slow drawdown of 27 by 1986 (3 years after EOC). During operation - - - - - settlement between the core, filters mostly in rockfill, but also
start of 1991. construction, and were
m from 1997 to mid 1998. piezometers show large drop in PWP across the and rockfill during construction. filters and core.
abandoned.
core.

Tokachi - no information - No information - 1000 - - - - - - -

During construction - PWP in the core were


measured at up to 70% of the applied
Started mid 1985, reached FSL by embankment load. Slow dissipation post
Upper Dam 0.75 to 1.2 years Fluctuating Rapid daily drawdowns of up to 20m. - - - - - - - -
Nov 1985. construction (still dissipating more than 8 yrs
later). Limited response of core to reservoir
fluctuation.
Appendix F Page F29

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 25 of 30)
SURFACE MONITORING - POST CONSTRUCTION
Crest Upstream Slope Downstream Slope
Time of Initial
Name Readings Comments
SLT (% per Comments Comments on Deformation and Observations.
(years after Settlement Displacement Settlement Displace- Settlement Displace-
log cycle of
EOC) (mm) (mm) (mm) ment (mm) (mm) ment (mm)
time)

0.1 (upstream edge


The majority of the post construction settlement and displacement
of crest) 329 or 0.44% 68 or 0.12% (19 127 mm Embankment used as pumped storage for peak power generation. Subject to
Split Yard Creek 0.10 31 mm (19 yrs) - - occurred during FF. Very low long-term settlement rate for the crest. Consider deformation behaviour as "normal".
3.93 all other (19 yrs) yrs) (19 yrs) rapid daily reservoir fluctuations.
IVM indicates uniform post construction settlement profile with depth.
SMPs.

Longitudinal cracking on crest observed due to greater magnitude


Large magnitude crest settlement on FF then steady long-term rate.
of settlement of upstream rockfill than core, most likely due to Upstream crest displacement on first filling is unusual and potentially "abnormal".
874 or 1.01% Greater settlement of upstream rockfill on FF. Displacement of crest
South Holson 0.08 to 0.13 0.47 -73 (6.6 years) - - - - collapse settlement of poor quality, dumped and sluiced rockfill. Accelerations of crest settlementon large drawdown in first four years, but not
(6.6 years) is upstream on FF, then trend is virtually steady (with fluctuations
Acceleration in crest settlement on draw-down at 3 and 4 yrs, but observed after this.
due to reservoir fluctuation).
not on large drawdown at 5 yrs.

Limited information. Not sure of the reason for the large First filling not completed within period of collected records, still more than 20 m
Limited data records. First filling not complete in mid 1981. Crest
857 mm or 800 mm or 0.8% magnitude settlement of the downstream shoulder, maybe it is due below FSL. The upstream rockfill zone between the cofferdam (crest 70 m below
Srinagarind 0.04 - - - - - settlement shows acceleration on rising reservoir level. Large
0.61% (2.9 yrs) (2.5 yrs) to collapse compression in the outer reasonably compacted rockfill main dam crest) and main dam core was flooded prior to the end of construction
settlement of downstream shoulder.
zone. to reduce the deformation on first filling.

Large downstream displacement likely to be associated with


Large downstream displacement and settlement of crest and upper Stored water prior to end of construction, to within 40 m of FSL, but not likely to
799 mm or 961 mm or 901 mm collapse compression in the downstream shoulder (post
Svartevann 0 0.87 1200 mm (4 yrs) - - slopes during latter stages of first filling (from 1.8 to 2.2 years). Long- have greatly affected the post-construction deformation behaviour. Consider
0.62% (4 yrs) 1.01% (4 yrs) (4 years) construction strains in the shoulder are high). Consider
term crest settlement rate is also high. deformation behaviour as potentially 'abnormal' on first filling.
deformation behaviour as potentially "abnormal".
no indication, 331 mm or
Several periods of acceleration in crest settlement in first 4 years, Constructed in broad river gully with 25 to 38 degree abutment slopes. Crest
Taisetsu suspect close to 0.38% (13.6 0.22 - - - - - Limited information.
thereafter near contant settlement rate (rate per log time). settlement considered normal.
EOC years)

Displacement of SMPs on the upper slopes and crest is downstream.


Long-term displacement rate is near zero for the crest and upstream Consider post construction deformation behaviour as 'normal'.
-0.2 for SMPs on
540 mm or 714 mm or 184 mm 512 mm or 305 mm shoulder, and very low for the downstream shoulder. Settlements ES1 (IVM) shows greater settlements in lower core, which is Consider deformation behaviour as 'normal'. Reservoir has remained close to
Talbingo upper slopes, 0 for 0.22 220 mm (28 yrs)
0.35% (28 yrs) 0.49% (28 yrs) (28 yrs) 0.35% (28 yrs) (28 yrs) show normal trend. Greater settlements of the upstream shoulder probably due to consolidation on dissipation of construction PWP FSL since first filled.
crest.
may be due to some collapse settlement in the upstream (Zone 3C) in the wet placed core material.
rockfill.

Stress distribution indicates significant arching across the core.


no indication, Only have crest settlement data which indicates typical settlement Consider post construction deformation behaviour as 'normal', although only
675 mm or HSGs in core during construction indicate possible zones of plastic
Tedorigawa suspect close to 0.27 - - - - - behaviour. Several periods of acceleration during the period of FF. have limited data. Possible development of plastic deformations in core during
0.44% (9 yrs) deformation of the wet placed core in mid to lower regions of core,
EOC The long-term rate is steady (in log time). construction.
with strains as high as 7 to 8%.

0.16 years after 263 mm or Relatively large settlement on first filling (185 mm), long-term rate is Constructed in broad river gully with 30 to 35 degree abutment slopes. Crest
Terauchi 0.17 - - - - - Limited information.
EOC 0.32% (13 years) near constant (in log time). settlement considered normal. Relatively flat slopes used in design.

Acceleration in settlement rate during periods of rising reservoir level


Consider as 'normal' deformation. Settlement plots indicate an Consider as case of 'normal' deformation behaviour. Well compacted core and
on FF. Net displacement on FF of the upper slopes and crest is
0 - SMPs on slopes 1060 mm or 1010 mm or 310 mm 735 mm or 540 mm acceleration of the crest settlement on drawdown in 1997-98, but rockfill shoulders. 300 mm greater settlement of upstream slope than
Thomson 0.21 155 mm (15 yrs) downstream, but the initial displacement of crest and upstream slope
0.48 - crest 0.64% (15 yrs) 0.75% (15 yrs) (15 yrs) 0.54% (15 yrs) (15 yrs) the magnitude of settlement on drawdown is small (less than 30 to downstream slope for SMPs at similar elevations. Difference likely due to some
was upstream. Relatively large magnitude of displacement of the
40 mm). collapse settlement of the upstream sedimentary rockfill on wetting.
crest and downstream shoulder on FF.

no indication,
160 mm or Relatively small magnitude crest settlement. No indication of start of Constructed in broad river gully with 41 to 54 degree abutment slopes. Crest
Tokachi suspect close to 0.17 - - - - - Limited information.
0.19% (6 yrs) initial monitoring. settlement considered normal. Relatively flat slopes used in design.
EOC

Longitudinal cracking on the crest (up to 130m length and 50 mm


830 mm or
Settlement of the downstream edge of the crest shows a normal width) first observed in Jan 1987. Filled with slurry grout. Further
0.94% (d/str Differential crest settlements and crest spreading resulted in cracking which has
trend, except that the long-term settlement rate is high. On FF cracking in Feb 1989, Feb 1991, Aug 1992 and 1993/94. Cracks
Upper Dam 0.17 years edge at 8 yrs), 0.75 - - - - - persisted for a period of 8 years. The persistence of the cracking is a possible
greater settlements of the upstream edge of the crest (cf located on the up and downstream side of core, possibly due to
greater sett of indicator of "abnormal" behaviour. Limited available information.
downstream edge) were measured and crest spreading. differential settlement as a result of collapse settlement of the
upstream edge.
rockfill.
Appendix F Page F30

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 26 of 30)
GENERAL DETAILS CONSTRUCTION / DESIGN
Embankment
Construction Timing Dimensions
Classification
Name Owner/ References
Location Geology Foundation Dam Core Upstream Downstream Comments on Design
Authority Year Time Core Height, Length, Ratio
Zoning Width / Slope Slope
Completed (years) n Type H (m) L (m) L/H
Class Slope (H to V) (H to V)

Broad central core supported by poorly compacted rockfill


shoulders. Core zone comprises inner central core of SMEC (1998b)
Melbourne Sedimentary -
Upper Yarra Australia, Victoria Rock 1957, Mar 3 5,0,0 c-tk CL 90 610 6.8 2.4H to 1V 2.4H to 1V medium plasticity clays and outer core zones of clayey sands MMBW (1986)
Water Corp. sandstone, siltstone
and gravels. Outer slopes flatten out from 2H to 1V in the Connell Wagner (1998d)
upper section to 3.5H to 1V in the toe regions.

Narrow central core supported by dry placed and reasonably


compacted rockfill shoulders. Dual filters up and
Vatnedalsvatn Norway - - Rock 1983, Sept 3.9 5,2,0 c-tn SM 121 482 4.0 1.5H to 1V 1.5H to 1V Myrvoll et al (1985)
downstream of the core. Dam has a slight arch to upstream.
For hydro-electric purposes.

Broad central high plasticity clay core supported by poorly Wood (1960)
Mighty River Core on bedrock, shoulders on alluvials
Waipapa New Zealand Volcanic - ignimbrites. 1960, May 0.8 5,1,1 c-tk CH/SC 36 183 5.1 3H to 1V 2.5H to 1V compacted (?) rockfill shoulders. Random fill zone on Bialostocki (1961)
Power (ignimbrite silts and gravels)
downstream shoulder. Downer ES (1998)

Broad central clay core (0.85H to 1V up and downstream) Leonard and Raine (1958)
Tennessee 1.1
Watuaga USA, Tennessee Sedimentary - shale Rock 1948, Dec 5,2,1 c-tk CL/SC 94 274 2.9 2H to 1V 2H to 1V with thin filters and dumped and sluiced rockfill shoulders. Blee and Meyer (1955)
Valley Authority (seasonal)
Outer slopes in series of benches sloped at 1.4H to 1V. Blee and Riegel (1951)

Central core supported by compacted rockfill shoulders. SMEC (1996b)


Goulburn
William Hovell Australia, Victoria Volcanics - rhyodacite Rock 1971, Apr 1.5 5,2,0 c-tm ML 34 357 10.5 1.33H to 1V 1.43H to 1V Steep slopes. Downstream slope 1.33H to 1V with a single Howley (1971)
Murray Water
bench. Cox (1972)

Central clay core supported by compacted rockfill shoulders.


Australia, New State Water Sedimentary - grit, Small cofferdam at upstream toe and random fill zone at Straw et al (1985)
Windemere Rock 1984, Apr 1.5 5,2,1 c-tm CL 67 825 12.3 1.65H to 1V 1.55H to 1V
South Wales DLWC sandstone and shale downstream toe. Core centreline slightly upstream of axis LWC NSW (1997)
(0.6H to 1V upstream, 0.2H to 1V downstream).

Core on bedrock. Shoulders on alluvial Narrow central clay core with well compacted rockfill
Australia, QDNR (now 6 QIWSC (1976)
Wivenhoe Sedimentary - sandstone sands and gravels (no indication of 1982, Mar 5,2,1 c-tn CL 59 1200 20.3 1.75H to 1V 1.75H to 1V shoulders. Gravels used in upper upstream shoulder. Upper
Queensland Sun Water) (staged) QWRC (1979a, 1979b)
strength or density) slopes at 1.5H to 1V and lower slopes at 2H to 1V.

Central core supported by compacted rockfill shoulders.


Existing concrete gravity dams forms the upstream toe of the
Australia, New State Water PWD NSW (1992b)
Wyangala porphyritic gneiss Rock 1968, June 2.5 to 3 5,2,0 c-tm SC-SM 86 1510 17.6 1.5H to 1V 1.6H to 1V embankment. Steep upper embankment slopes (1.3H to 1V
South Wales DLWC LWC NSW (1994)
for upper 6 m). Dual filter zones up and downstream of the
core.
Appendix F Page F31

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 27 of 30)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Relative
Layer
ASCS Ratio or Dry Moisture ASCS Density or Level of
Zone Source n Placement Methods Comment Zone Source n Thickness Plant Comment
Class Density Content Class Dry Density Compaction
3 3 (mm)
(t/m ) (t/m )

Sandy clays of medium plasticity for the


Spec. 95 -
CL (CH in upper Well compacted - 200 mm placed dry of OMC most part. High plasticity clays used in No specific filters used. Fine rockfill used at
1 (inner clay 105% SMDD
Upper Yarra alluvial, colluvial third) - sandy (loose) layers, 12p sheepsfoot (Spec. 2% dry of the upper third of the core. Zone 1 is the - - - - - - - some locations, either up or downstream of
core) (bulk density =
clays 3 and grooved drum rollers OMC to OMC) inner core zone and its geometry varies coarse rolled fill sections.
2.05 t/m )
with depth.

Sandy gravels with < 2% non-plastic fines. Dual


SM - gravelly 102.5% SMDD Well compacted - 500 mm Gravelly silty sands, max. size 150 mm, 2A & 2B alluvium (2A) & 2p 11t SDVR, no water reasonable to
Vatnedalsvatn 1 morraine 3 0.5% wet of OMC GW/GP - 1000 filters of thin width up and downstream of the
silty sands 2.09 t/m layers, 6p 11t SDVR 31% fines. (filters) crushed rock (2B) added good
core.

Well compacted filter zones. Zone 2 upstream


Well compacted - 200 mm Broad core, 1H to 1V upstream and 0.5H
98.5% SMDD GW/GP - sandy reasonable to of core (3 m width) and Zone 2A downstream of
Waipapa 1 Ignimbrite ('soft') CH/SC 3 (loose) layers, 12p sheepsfoot 1.4% dry of OMC to 1V downstream. Some weaving 2 & 2A (filters) Ignimbrite rockfill DR = 70% 450 4-9p sheepsfoot rollers
1.0 t/m gravels good core (3m width). Relatively coarse (max. >
rollers problems when wet.
150mm), 6% fines.

Well compacted (?) - 150 to


possibly placed close Low plasticity sandy clays to clayey 2A & 2B limited information. Dual filters of thin width (6.1
Watuaga 1 - CL/SC - 200 mm layers, sheepsfoot quarried rock ? - - - -
to or wet of OMC sands. (filters) m wide) up and downstream of the core.
rollers

ML (MH/CL) - Well compacted - 150 mm Medium to high plasticity sandy 2A & 2B alluvial sands and SP (2A) Dual filters up and downstream of the core, 2.7
William Hovell 1 colluvial 99.4% SMDD 0.4% wet of OMC Spec > 70% DR < 450 SDVR (good)
sandy silts/clays layers silts/clays. (filters) gravels GP (2B) to 4.2 m combined width.

Medium plasticity (but varies from low to High density ratio indicates the filters are well
CL (SC, CH,
residual andesite & Well compacted - 150 mm high) gravelly sandy clays. Average of 2, 2A & 2B alluvial sands and 2A - SP compacted. Dual filter (2A & 2B) downstream of
Windemere 1 GC) - gravelly 101.5% SMDD 0.6% dry of OMC D R =85% 400 track rolled by dozers. Good
colluvium layers, Cat 825 padfoot roller 60% fines. 0.6 m contact layer placed (filters) gravels 2, 2B - GP core of 5 m width, single filer (Zone 2) of 2.5 m
sandy clays
slightly wet of OMC. width upstream of the core.

CL - sandy 2A - SW Very high relative densities achieved in filters.


clays, 83% Well compacted - 150 mm Low to medium plasticity sandy clays. 2A, 2B & 2C 2B -GP DR ~ 90% (85 to Dual filters used up and downstream of the core.
Wivenhoe 1 alluvial 101.1% SMDD 0.2% wet alluvial (washed) 300 SDVR, water added. Good
average fines layers, padfoot rollers 0.6 m thickness wet placed contact layer. (filters) 2C - GP to 156%) Combined width of about 3 m. Filters under
content SC/SM downstream shoulder.

Well-compacted filter zones. Combined width of


alluvial sands (2A)
SC/SM - silty to 102% SMDD Well compacted - 150 mm Well compacted central core of clayey to 2A & 2B SW/GW (2A) Spec. >70% DR 5 to 6.7m. Additional rockfill transition zone
Wyangala 1 Residual granite 3 1% dry of OMC & crushed rock 3 300 8.5t SDVR Good
clayey sands 2.02 t/m layers, 8p sheepsfoot roller silty sands. (filters) GW (2B) 1.87 to 2.02 t/m (Zone 3A) of 3.4 m width used on downstream
(2B)
side of core.
Appendix F Page F32

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 28 of 30)
MATERIALS
Rockfill Zone Rockfill Zone (in some cases earthfill stabilising berm)
Name
Dry Layer Level of Dry Layer Level of
Grading Clean / Grading Clean /
Zone Source Density Thickness Placement Method Comp- Comment Zone Source Density Thickness Placement Method Comp- Comments
Description Dirty 3 Description Dirty 3
(t/m ) (m) action (t/m ) (m) action

Well compacted clayey sands to clayey


12p sheepsfoot and
gravels with low plasticity fines. Zone 1A
sandstone & indication of high Spread in 1.5 m layers, decomposed GC/SC - clayey grooved drum rollers,
Poor compaction of weathered rockfill. 1A (outer (bulk density used up and downstream of the central clay
Upper Yarra 3 siltstone, HW to fines content for dirty (?) - 1.5 no indication if water poor sandstone and sands to clayey - 3 0.2 m moisture specification is good
Suspect breakdown under trafficking likely. core zone) = 2.08 t/m ) core. Of variable geometry. Outer slopes of
MW some rockfill added. siltstone gravels. from 2% dry of OMC to
Zone 1A are approx 1H to 1V. Average
OMC
fines = 29% (14% clay fraction).

quartz, granite 6p 11t SDVR, no water Placed without addition of water. No


Vatnedalsvatn 3 - - - 1.5 Reasonable - - - - - - - - -
and gneiss added indication of grading.

Not formally compacted and placed without


very poor
Not formally compacted. water addition. Some breakdown on sands and 6m (rockfill) Rockfill dumped, Sands
max up to 1.2m (?), 1.8 (3A) (rockfill), good Random fill zone on downstream slope.
Waipapa 3A & 3B ignimbrite ('hard') - - No indication that water Poor handling. Zone 3A used in the downstream 4 gravels, and - - - 0.2m sands & & gravels 10-12p 32t
variable. 0.9 (3B) (sands & Underlain by foundation drain.
added. shoulder and Zone 3B used in the upstream ignimbrite gravels pneumatic tyred rollers.
gravels)
shoulder.

Sound, durable rock of at least high


Watuaga 3 quartzite - - - 9.1 dumped and sluiced poor - - - - - - - - -
unconfined compressive strength.

Rockfill mostly of 'hard' fresh Rhyodacite,


max 1.2 to 1.5m 3
clean (some 1.93 t/m but some softer rock. Breakdown of
William Hovell 3 rhyodacite 13% < 19 mm 1.5 8p 4.5t SDVR reasonable - - - - - - - - -
dirty zones) D R = 75% surface of layer (upper 150 mm) on
2 - 4% fines
placement.

Very high strength rock. Used as the inner 2p 10t SDVR (3B) Very high strength rock. Zone 3B used
3 3
andesite (very (2.15 t/m - 4p 10t SDVR, no reasonable to rockfill zone upstream and downstream of andesite (very (2.15 t/m - 6p 10t SDVR (3C) predominantly as outer rockfill zone. Zone
Windemere 3A - clean 1.2 3B & 3C - clean 2.0 reasonable
high strength, FR) bulk density) indication if water added. good the core. Narrow rockfill transition high strength, FR) bulk density) no indication if water 3C used as outer rockfill in downstream
downstream of filters. added. shoulder below about mid height.

3
1.98 t/m 4p 10t SDVR, water MW to FR rock used. FR is high strength,
3B up to 1m max. 3 Well-compacted gravels as indicated by very
sandstone (ripped dirty (3B 98% SMDD for 3B - 1.0 added (slightly dry of MW to SW is medium to high strength. 3A GP/GW - sandy 2.20 t/m 2p (?) 10t SDVR, water
Wivenhoe 3B & 3C 3C up to 0.3 m Good alluvial (?) clean < 1.0 Good high relative densities achieved. Zone 3A
& blasted) definitely) weathered 3C - 0.5 OMC for weathered Zone 3B is the main rockfill. Zone 3C used (gravels) gravel D R = 110% added
max. used in upper upstream shoulder.
rockfill rockfill) for up and downstream cofferdams.

High to very high strength porphyritic gneiss High to very high strength porphyritic gneiss
2.00 to 2.24 minimum 3p 8.5t SDVR, Reasonable to (UCS = 55 to 160 MPa). Zone 3B used in 2.00 to 2.24 6p 8.5t SDVR, no water Reasonable to (UCS = 55 to 160 MPa). Zone 3C used in
Wyangala 3B porphyritic gneiss - Clean 3 1.2 3C porphyritic gneiss - Clean 3 2.4
t/m no water added good the inner shoulder regions up and t/m added poor the outer shoulder regions up and
downstream. downstream.
Appendix F Page F33

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 29 of 30)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Settlement
Pore Pressures in Core (Average Rockfill and/or Filters Comment Internal Monitoring
(from end of Steady / Instrument
Comments Comment mm/year) Comment Equipment
embankment Fluctuating Type/No.
mm % of H
construction)

Strain profile shows normal


Annual seasonal drawdown from 5 to During operation - steep phreatic surface across
trend. Strains increase with
30 m. Large events of 27 to 33 m in - core. Upstream Zone 1A fluctuates with reservoir IVM, 2 (in
2380 (ES1) & 2.69% (ES1), depth below the crest, 0 to 1%
Upper Yarra 0.41 to 1.78 years July 1957 to Dec 1958 Fluctuating 1966/68 (33 m), 1976, 1979, 1981/83 the level but at a reduced amplitude. Some to - central Zone 1 - - -
2190 (ES2) 2.64% (ES2) in upper 10 m increasing to 3.5
(33m), 1990 to 93 (23 to 30m negligible response in central clay core and core)
to 4.5% near base. Their is
annually). downstream Zone 1A.
some variability.

Jan 1981 to Dec 1983. Water stored Negligible positive pwp during construction. PWP
Vatnedalsvatn -2.75 to 0.17 years during construction, to within 10 m of Fluctuating Seasonal drawdown of 40 to 60m. fluctuates with reservoir level but a lower - - - - - - - -
FSL before EOC. amplitude.

Waipapa - Unknown. Possibly in 1963. Steady Steady at FSL. - - - - - - - - -

Dec 1948 to Sept 1949 (within 3 m of annual drawdown, to lowest level of


Watuaga 0 to 0.83 years Fluctuating - - - - - - - No monitoring during construction -
FSL) 24 m below FSL (to 1957).

Annual drawdown of 6 to 18 m.
William Hovell 0 to 0.6 years April to Sept 1971 Fluctuating Largest drawdowns (to 18 m) in 1991 - - - - - - - - -
and 1998 (20 and 27 after EOC).

Plot of vertical strain versus


During construction - limited positive PWP in core, IVM, 1 (7 m
slow rise in reservoir level to FSL depth shows some variability.
Windemere 0 to 6 years Steady - max r u of about 0.1. - upstream of 2134 3.2%. - - -
from 1984 to 1990. Vertical strain range from 2 to
During operation - steady. dam axis)
6%.

During construction - some piezos in core showed Rockfill - 255 mm or 1.4% IVM results indicate significant
1983 to 1988. Possibly completed Not annual drawdown. Subject to Vertical strain increases with
Fluctuating positive PWPs (ru up to 0.2, possibly higher) with 1085 (ES1) (0.8 to 2.0%). Does not settlements occurred within the
Wivenhoe 0 to 6.4 years earlier, but have no records between slow fluctuation, no rapid drawdown - IVM - 2 in core 1.55 to 1.85% increasing fill depth from 0% INs in core and shoulders.
(slow) slow dissipation post construction. Others in core 840 (ES4) include deformation in foundation under the shoulders (415
3.6 and 5.6 years. events. near crest to 4 - 5% at depth.
showed negligible to limited positive PWP. alluvial foundation. mm to Dec 1989).

Vertical strain profiles indicate


Existing concrete gravity dam at Generally within 5 - 10 m of FSL. HSGs & IVMs indicate settlements
1530 (A) 1.83% (A) normal behaviour. Strains IHMs and HSGs in
upstream toe, stored water during Fluctuating Low period 1979 to 1984 (11 to 15 yrs of similar magnitude in the core and
Wyangala -1.5 to 5.7 years. - - IVM - 3 in core 1550 (B) 1.87% (B) range from 0.5 to 3% for the - downstream section (at main
construction. Reached FSL by Feb (slow) after EOC) when the reservoir was filters zones, and slightly greater in
572 (C.) 1.61% (C.) most part, and 4 to 4.5% in the section).
1974. gradually drawn down some 45 m. the rockfill during construction.
base of the core.
Appendix F Page F34

Table F1.1: Central core earth and rockfill embankments case study information (Sheet 30 of 30)
SURFACE MONITORING - POST CONSTRUCTION
Crest Upstream Slope Downstream Slope
Time of Initial
Name Readings Comments
SLT (% per Comments Comments on Deformation and Observations.
(years after Settlement Displacement Settlement Displace- Settlement Displace-
log cycle of
EOC) (mm) (mm) (mm) ment (mm) (mm) ment (mm)
time)

Deformation indicates 'abnormal' behaviour during large


Initial downstream displacement on FF of the crest and upper slope drawdown. During the 1966-68 drawdown the crest settled up to
regions. The crest displacement shows a trend to upstream at 9 100 mm at the main section and IVM ES1 (installed in the core)
0.38 years (prior to 533 mm or 425 mm or 37mm 398 mm or 390 mm "Abnormal" deformation behaviour during large drawdown events, in particular on
Upper Yarra 0.57 140 mm (41yrs) years, then a steady to slightly downstream displacement after about shows that most of this (about 75 mm) was concentrated at depths
start of ff) 0.60% (41yrs) 0.58% (41yrs) (34yrs) 0.53% (41yrs) (41yrs) the first large drawdown in 1966/68.
27 years. Acceleration in the crest settlement rate during the large below crest level of 11.2 to 14.3 m. On subsequent large
drawdown events of 1966-68, 1979 and 1993. drawdowns, the magnitude of crest settlement decreased; 30 to 35
mm in 1979/80 and 22 mm in 1993.

400 mm Storage of water prior to EOC likely to have affected monitored Stored water to close to FSL before end of construction (within 10 m of FSL).
Large downstream displacement of crest and upper slopes during
400 mm or 400 mm or (0.75 yrs), 180 mm or post construction deformation due to the susceptibility of the dry Likely to have affected the post-construction deformation behaviour.
Vatnedalsvatn 0 - 200 mm (0.75 yrs) 250 mm last 10 m of filling to FSL. Records only until 0.75 years after end of
0.33% (0.75 yrs) 0.37% (0.75 yrs) questionable 0.17% (0.75 yrs) placed and reasonably compacted rockfill to collapse compression Downstream displacements during the latter stages of first filling (I.e. last 10 m to
construction.
reading ? on wetting. FSL) were of large magnitude.

Limited records for early years of operation. Very low long-term Consider as 'normal' deformation behaviour. Suspect that even
0.79 years (before 36 mm or 0.10% 39 mm or 0.13% 23 mm settlement rate. Displacements of the crest and downstream though water was not added to the rockfill during placement, the Consider as 'normal' deformation. Relatively small embankment (H = 36m) and
Waipapa 0.075 41 mm (38 yrs) - -
start of ff) (38 yrs) (38 yrs) (38 yrs) shoulder are downstream, and the rate is virtually zero long-term wet climate provided sufficient wetting such that collapse deformations are relatively small.
(from 25 to 38 yrs). compression on first filling was not an issue.

SMPs in the centre of the crest. Greater magnitude settlement of the Longitudinal cracking on crest, coincident with the up and
0.08 (1 month after 970 or 1.03% Suspect "normal" behaviour for this type of embankment. Possibly some
Watuaga 0.44 - - - - - up and downstream shoulders than the core, 150 to 200 mm greater. downstream edges of the core. Due to greater settlement of the
EOC) (8.2 years) collapse settlement of the upstream rockfill on first filling.
No significant displacement on first filling. rockfill shoulders from collapse compression ?.

Crest settlement (in centre) shows acceleration in the settlement rate


Consider as 'normal' for the most part. Acceleration in crest Overall, consider the deformation behaviour as 'normal'. Potential abnormalities
on the first 2 drawdowns greater than 9 m, thereafter the settlement
86 mm or 0.24% 131 mm or 77 mm or 0.33% 67 mm settlement on the first 2 to 3 drawdowns and the relatively high possibly associated with steep slopes and "highly stressed conditions", and
William Hovell 0.39 years 0.15 38 mm (28 yrs) - rate is steady. Increased settlement rate for upstream shoulder after
(28 yrs) 0.47% (28 yrs) (28 yrs) (28 yrs) long-term settlement rate of the upstream shoulder are indications some collapse compression in the rockfill on wetting and reduced lateral support
5 years. Crest displacement shows upstream deformation during first
of potentially "abnormal" deformation behaviour. for the core.
3 drawdowns then gradual, very slow downstream displacement.

Consider as 'normal' deformation behaviour. The zone of relatively


Consider as 'normal' deformation. Possible broad shear zone in core at close to
Settlement profiles show normal trend. Displacement of the crest high vertical strain in the core is close to the upstream core / filter
405 mm or 252 mm or 45 mm 231 mm or 212 mm its upstream interface in the upper region of the core. In this upper region, Zone
Windemere 0.13 years 0.36 90 mm (14 yrs) and shoulders is downstream. IVM shows zone of high strain post interface and it occurred on FF. It is probably associated with
0.60% (14 yrs) 0.43% (14 yrs) (14 yrs) 0.40% (14 yrs) (14 yrs) 3B is directly upstream of the filters and is considered more susceptible to
construction from 10 to 30 m depth below the crest. greater settlement of the upstream shoulder on FF, partly due to
collapse compression than the reasonably to well compacted Zone 3A.
collapse compression.

Settlement show normal trends. Interrnal post construction


0 yrs (crest sett) No indication of any problems with embankment behaviour from
181 mm or 38 mm or 0.08% 59 mm 75 mm or 0.32% 69 mm settlement in the core shows uniform profile. Displacement of the
Wivenhoe 1.94 yrs (slopes & 0.20 65 mm (16 yrs) surveillance reports. Consider deformation behavioour as Consider deformation behaviour as 'normal'.
0.32% (16 yrs) (16 yrs) (16 yrs) (16 yrs) (16 yrs) crest and upper shoulders (both down and upstream) is downstream
crest displ) "normal".
on and post FF.

Fine cracks (transverse and longitudinal) observed in crest prior to


Displacement monitoring only captured the last 10 m of FF. The
1994. Reference to sink holes and settlements in crest (LWC
downstream crest and slope displace to downstream, and the
-186 mm (31 yrs) - NSW 1994) but no details. IVM B (upstream of axis and in core) Deformation behaviour 'normal' except for broad "shear zone" in the core near to
0.39 - crest sett upstream crest and slope displace to upstream. Crest spreading =
376 mm or u/str edge 419 mm or -71 mm 339 mm or 125 mm indicates large strains localised to 22 to 30 m depth in the core its upstream interface caused by differential settlement between the core and
Wyangala 0.64 - slope sett 0.18 80 to 90 mm on FF, and reached close to 200 mm by 1999 (31 yrs).
0.44% (31 yrs) 30 mm (31 yrs) - 0.53% (31 yrs) (31 yrs) 0.43% (31 yrs) (31 yrs) close to its upstream interface. Possibly indicates a zone of upstream shoulder. Acceleration in settlement of the crest and upper slopes on
3.8 - displ Acceleration in settlement of the crest and upstream slope on
d/str edge plastic deformation or a broad shear zone within which first drawdown.
drawdown from 1979 to 1981, and again in 1994/95 (although of
deformations were initially triggered on first filling, but later on
much lower magnitude).
drawdown of the reservoir.
Appendix F Page F35

Table F1.2: Zoned earth and rockfill embankments case study information
GENERAL DETAILS CONSTRUCTION / DESIGN
Construction Timing Embankment Classification Dimensions
Name Owner/ References
Location Geology Foundation Dam Upstream Downstream Comments on Design
Authority Year Time Core Core Height, H Length, Ratio
Zoning Slope Slope
Completed (years) n Width Type (m) L (m) L/H
Class (H to V) (H to V)

Core and downstream shoulder on Narrow central core with shoulders of decomposed granite Kim (1979)
Andong Korea - Granite (?) bedrock, upstream shoulder on 1976, June 4 4,1,1 c-tn SM 83 542 6.5 2H to 1V 1.7H to 1V and rockfill. Cofferdam forms upstream embankment toe. Sonu (1985)
decomposed granite. Filters encase downstream decomposed granite zone. ADCO (1976)

In the floodplain, most of core and Broad central core (1.7H to 1V upstream slope and 0.5H to
shoulders are founded on alluvial 1V downstream) supported by outer earth and rockfill USBR (1959)
Bradbury USA, California USBR Sedimentary - siltstone & shale 1953, Feb 1.5 4,1,1 c-tk SC/GC 85 1022 12.0 3.25H to 1V 2.25H to 1V
sands and gravels (up to 20 - 22 m zones. Broad cut-off trench to bedrock, located upstream of Stateler (1983)
depth). Cut-off trench to bedrock. dam axis.

Broad central core with filters and outer dumped rockfill Douglas (1965)
Sedimentary (low grade Core and abutments on bedrock.
Australia, New State Water shoulders. Cofferdam incorporated into upstream toe. WCIC NSW (1964)
Burrendong metamorphism) - sandstone & Rockfill shoulders in river section 1963, Oct 2.25 4,2,1 c-tk CL 76 1130 14.9 2.5H to 1V 2.25H to 1V
South Wales DLWC Core comprises central clay core with up and downstream PWD NSW (1992a)
greywacke on gravels.
zones of sandy silts and clayey sands and gravels. WRC NSW (1992)

7 (staged)
Core and broad transition zones on Narrow central clay core with broad transition zones and Alvarez & Bravo (1976)
Sedimentary - sandstones of 1979-81 to
sandstone bedrock. Steep shoulders of limestone rockfill. Cofferdam at upstream toe. Bravo (1979)
Canales Spain - Tertiary age. Strongly 1986, Sept 100m 4,1,0 c-tn CH 156 378 2.4 1.75H to 1V 1.7H to 1V
abutments on bedrock. Shoulders Constructed in broad river gully (100m wide) with near Giron (1997)
tectonized. 1985-86 to
in river section on gravels. vertical abutment slopes. Bravo et al (1994)
156m

Core zone comprises an inner Zone 1 zone (SC/CL)


Melbourne SMEC (1971)
supported by broad silty sand zones (Zone 1A). Outer
Cardinia Australia, Victoria Water granodiorite Rock 1973, May 1.7 4,2,1 c-tk SC/CL 86 1542 17.9 1.85H to 1V 2H to 1V MMBW (1987a)
downstream earthfill shoulder and upstream rockfill
Corporation Connell Wagner (1998a)
shoulder.

Soil foundation. Low plasticity


sandy clays, clayey sands, clayey
Very broad central clay core. Thin outer shoulder zones of
Sedimentary - sandstone & gravels and sandy silts. Up to 10
Dixon Canyon USA, Colorado USBR 1948, late (?) 1 (approx) 4,1,0 c-vb CL 74 387 5.2 2.5H to 1V 2.5H to 1V rock fines and rockfill. Broad (55 m width) cut-off to USBR (1946, 1997)
shale m depth in valley and up to 5 m
bedrock, aligned upstream of dam axis.
depth on abutments. Cut-off to
bedrock.

SMHEA (1998a)
Meta-sediments, low grade - Very broad central core (1.6H to 1V up and 1H to 1V
Australia, New SMEC (1990)
Eucembene SMHEA quartzite & hornfels; with Rock 1958, May 2.5 4,1,0 c-vb SC 116 579 5.0 2.5 - 3.5H to 1V 2.0 - 3.5H to1V downstream) of clayey sands with outer shoulders of sands
South Wales USBR (1953)
andesitic dyke intrusions. and gravels, and rockfill. Slopes flatten toward toe.
DPW NSW (1955)

Melbourne Basalt overlying granodiorite. Broad central core with downstream earthfill zone and MMBW (1972, 1988)
Variable - basalt and completely to
Greenvale Australia, Victoria Water Layer of sediments between 1970, Aug 1.5 4,2,2 c-tk SM 52 2500 48.1 1.8H to 1V 3H to 1V upstream rockfill zone. Bench on upstream slope at 18 m Reinhold (1969)
highly weathered granodiorite.
Corporation geological units. below crest, below this the slope is 5H to 1V. Connell Wagner (1998b)

CFE Moderately karstic foundation.


Complex design. Very narrow central clay core with Ramirez de Arellano & Gomez
(Federal Sedimentary - limestone with Core on bedrock. Outer shoulders
La Angostura Mexico 1973, Nov 2.5 4,0,0 c-tn CL 146 295 2.0 2H to 1V 1.8H to 1V shoulders of sands and gravels, and rockfill. Cofferdam (1972)
Electricity shale seams on alluvium in river section (~ 15 to
forms upstream toe of embankment. Benassini et al (1976)
Commission) 20 m).

Narrow central core with shoulders of compacted sandy


Hydro Quebec Verma et al (1985)
2.5 gravels and rockfill. Rockfill used in the outer slopes and at
La Grande 4 (LG4) Canada (James Bay Granite and gneiss Rock. 1981, Nov 4,1,0 c-tm SM 125 3800 30.4 1.7H to 1V 1.7H to 1V Pare et al (1984)
(seasonal) the downstream toe. Cofferdam at upstream toe.
Project) McConnell et al (1982)
Constructed in broad glaciated valley.

Narrow central core with shoulders of sand to boulders.


Borovoi et al (1982)
Sedimentary - sandstones and Core on bedrock. Shoulders on Rockfill on outer face of slopes and downstream toe.
Nurek Russia - 1980 (?) 2 4,2,0 c-tn GC/GM 289 700 2.4 2.25H to 1V 2.1H to 1V Sokolov et al (1985)
siltstones alluvium in gully (approx 7 m depth) Cofferdam at upstream toe. Constructed in narrow, steep
Borovoi & Mikhailov (1978)
sided valley.
Appendix F Page F36

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 2 of 10)
MATERIALS
Earthfill Core (Zone 1) Filter or Transition Zones
Name Dry Density Relative
ASCS 3 Moisture Layer
(t/m ) or Density or
Zone Source n Placement Method Content at Comments Zone Source ASCS Classn Dry Density Thickness Placement Method Comment
Class Density Ratio Placement 3 (mm)
(% SMDD) (t/m )

99.7% SMDD Well compacted - 200 mm Silty sandy gravel with non-plastic fines. Well-compacted filters. Located upstream and
decomposed alluvial sandy DR = 82% 4p 10t SDVR, in moist
Andong 1 SM 3 layers, 8p 25t tamping 0.4% dry of OMC Max size 10 mm, 18% fines. Difficulties 2 GW 300 downstream of central earthfill and encase
granite 1.95 t/m roller with breakdown during compaction. gravels 1.87 t/m 3 condition
downstream decomposed granite (Zone 3)

Clayey sands and gravels of low fines


Well compacted - 150 mm GW/GP - sandy Well compacted. Zone 2 used as upstream
98.9% SMDD 0.9% dry of OMC content and low plasticity. Broad core DR = 89%
Bradbury 1 alluvial SC/GC layers, 12p (minimum)
1.81 t/m
3
(12.6 to 14.6%) zone with relatively flat upstream slope 2 (transition) alluvial gravel with cobbles,
2.09 t/m 3
300 to 450 4p 20t crawler tractors. shoulder earthfill and also transition zone
sheepsfoot roller < 5% fines downstream of the core.
(1.7H to 1V).

Low to medium plasticity gravelly sandy Good compaction of filters. Used up and
Well compacted - 230 mm 4p tamping rollers (2A)
(Spec. >98% (Spec. 3% dry of clays. Zone 1A is the narrow, central 2A - GC/GW 2A - 225 downstream of broad central earthfill zone
Burrendong 1A alluvial CL layers (loose), 12p 2 & 2A river gravels (Spec. DR >70%) 2-4p 4.5t SDVR (2),
SMDD) OMC to OMC) region of the core (0.25H to 1V slopes up 2 - GW/GP 2 - 300 to 600 (Zones 1A, 1B & 1C). Width narrow at crest
sheepsfoot rollers watered as required
and downstream). and broad at base (> 30 m).

Kalkirita
Limited information. Very broad transition
Spec. OMC to 2% Likely that core was reasonably to well (limestone
Canales 1 residual CH - - zones of clayey or silty gravelly sands.
wet of OMC. compacted. Silty clays of high plasticity. 2 (transition) reduced to sand SC/SM (?) - - -
Increasing width with depth, up to 25m wide
by tectonic
upstream and 60m wide downstream.
action)

1 - colluvial & 1 - (OMC to 2% Central core (Zone 1) of clayey sands to


Well compacted - 150 mm limited information. Filter zone of thin width
residual 1 - SC/CL (?) (Spec. >98% wet) sandy clays of medium plasticity. Outer
Cardinia 1 & 1A layers, 8p sheepsfoot 2B (filter) - GP/GW (?) (Spec. DR >70%) 450 2p 8t SDVR used up and downstream of the central earthfill
1A - residual 1A - SM SMDD) 1A - (1.5% dry to core zone (Zone 1A) of silty sands. 0.6
rollers zone (Zones 1 and 1A).
granodiorite 0.5% wet) m contact of CH clay placed wet of OMC.

Well compacted. Zone 2 located downstream


Mainly sandy clays of low plasticity, 2 - rockfill (rock GW/GP - 150mm
98.5% SMDD Well compacted - 150 mm Avg 2.8% dry of of central core. Zone 2A located upstream
Dixon Canyon 1 alluvial CL placed well dry of Standard Proctor 2 & 2A fines) max 10p tamping rollers, wetted
- 300 (upper 43 m) of central core. Width of Zone 2
1.78 t/m 3 layers, 12p tamping rollers OMC (transitions) 2A - sand and 2 - 17% < 6.75 mm, for compacteion
OMC. increasing with depth (up to 30 m wide at base
gravel 2A - 39% < 6.75 mm
of dam).

Clayey sand with low plasticity fines. Likely reasonable level of compaction. Used as
Well compacted - 150 mm river sand &
decomposed (Spec. >95% Spec. OMC to 2% Well compacted, placed on the dry side 2 and 4 GP/GW, high gravel Spread by dozer, 4p 18t transition (together with Zone 1A) from core to
Eucembene 1 SC layers, 12p (minimum) 18t gravel (4) and - 305
granite MMDD) dry of OMC. of OMC. Up to 20% of results outside (transition) content (> 70%) crawler tractors, watered outer rockfill both up and downstream of central
sheepsfoot roller. crushed rock (2)
the moisture spec. core.

Well compacted - 150 mm


Well compacted clayey silty sands to Well-compacted filters. Located upstream and
residual SM layers (225 mm loose), 4p 2A & 2B SP/SW (2A) 300 to 450
Greenvale 1 100.2% SMDD 0.3% wet of OMC sandy silts with medium to high plasticity crushed rock Spec. > 70% DR 4p 10t SDVR downstream of central earthfill and on
granodiorite (SC/ML/MH) Cat 834 padfoot and 4p (filters) GP/GW (2B) loose
fines. Fines content = 30 to 60%. downstream foundation.
groove roller

Well-compacted silty clay of medium


Well compacted - 250 mm Spec. OMC to 2%
La Angostura 1 - CL - plasticity. High plasticity clays used in - - - - - - No filters used.
layers, 8p 25t sheepsfoot wet of OMC.
upper 20 m of core.

Well compacted - 450 mm Well compacted gravelly silty sand. 25 Well compacted sandy gravels, < 5% fines.
La Grande 4 Spec. 1% dry to 2% select sand and High moduli. Zone 2A used up and
1 morraine SM Spec. > 97% SMDD (loose) layers, 4p 45t to 35% non-plastic fines. High modulus 2A (filter) GP/GW DR = 85% 450 (loose) 3p 5.5t SDVR
(LG4) wet of OMC gravel downstream of core, width narrow at crest
pneumatic tyred rollers. (approx. 250 MPa).
increasing to 12 to 15 m at depth.

Well compacted - 250 to moisture alluvium and


colluvium and Well compacted sandy clayey gravel. 2A & 2B colluvium SP/SW (2A) Well-compacted filters. Located upstream and
Nurek 1 GC/GM - 300 mm layers, 70t rollers conditioned prior to - 250 to 300 70t rollers (pneumatic ?)
weathered rock Max size of 200 mm. (filters) (screened and GP/GW (2B) downstream of central core. 10 to 16 m width.
(pneumatic ?) placement.
crushed)
Appendix F Page F37

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 3 of 10)
MATERIALS
Either Earthfill or Rockfill Zone Either Earthfill or Rockfill
Name
Dry Layer Layer
Grading Grading Dry Density
Zone Source Density Thickness Placement Method Comments Zone Source 3 Thickness Placement Method Comments
Description Description (t/m )
(t/m3) (mm) (mm)

4p tractor, no indication if
SM - silty gravelly
Well-compacted silty gravelly sands. Used up water added, but moisture Poor compaction. Limited information.
decomposed sand. Max size 10 to 97.9% SMDD 8p 25t tamping roller,
Andong 3 (earthfill) 3 300 and downstream of the central core. Similar to 4 (rockfill) rockfill - - 1000 content of 1% would Placed in outer downstream and
granite 15 mm, 13 - 15% 1.93 t/m 0.6% dry of OMC
central core material. suggest water was not upstream shoulders.
fines.
added.

Max 0.3 to 0.45m size,


12p sheepsfoot rollers, Well-compacted. Placed in the mid-section of
70% > 4.75 mm, cobbles and boulders (> Dumped and levelled off by Oversize from Zone 1 screening. Used
Bradbury 3 (rockfill) shales, weathered - 450 water added during the downstream shoulder. Broke down under 4 (gravels) alluvial - 900
some silt fines 75 mm) dozers. in downstream toe of embankment.
placement. trafficking by haul trucks and on spreading.
(medium plasticity)

Placed, spread and 1B - GC/SC, 15 to 40% Well compacted. Zone 1C located


Refered to as 'good quality' rock. Used as 16p sheepsfoot rollers
(2.05 t/m 3 bulk watered. High water 1B & 1C fines (Spec. > 96 to downstream and Zone 1B upstream of of
Burrendong 3 (rockfill) Greywacke - 1800 outer rockfill zone up and downstream and in alluvial 230 (loose) (1B), pneumatic rollers also
density) volume used but not high (earthfill) 1C - ML, 50 to 90% silt 98% SMDD) central Zone 1A . Together, 1A, 1B and
main body of upstream cofferdam. used.
pressure sluiced. fines. 1C formed central earthfill.

Zone 3B used in the outer up and downstream


shoulders. Reference to rock being hard and Thin (3m width) inner rockfill transition
limestone or coarse
Canales 3B (rockfill) limestone - - - - durable. Suspect either poorly compacted 3A (rockfill) - - - - zone placed up and downstream of
kalkirite
and/or dry placed due to large deformations on sandy transition zone.
FF.

Zone 3A (reasonably compacted) used in


3A & 3C Granodiorite, SW 3A - 1800 mm 4p 8t SDVR, no indication upstream outer shoulder. Zone 3C (well granodiorite, residual 4p 8t SDVR, no indication Used in downstream shoulder.
Cardinia - - 4 (rockfill) variable - 600
(rockfill) to FR 3C - 900 mm if water added compacted) foundation drainage layer soil to fresh rock if moisture added. Underlain by drainage layer of Zone 3C.
downstream of central core zone.

Sedimentary - not Not formally compacted. No indication of rock


placed and spread, no
Dixon Canyon 3 (rockfill) specific on which gravel to boulder size - 900 type or strength. Zone 3 used as the outer fill - - - - - - -
water added.
type zone in the up and downstream shoulders.

Difficulty meeting compaction (partic.


Spread by dozers and weathered granite GM/SM - variable gravel 12p (minimum) 18t
Very high strength rock. Zone 3 rockfill used (Spec. >95% when gravelly) and moisture
Eucembene 3 (rockfill) quartzite max 0.9 m - 900 sluiced (2 to 1 by volume 1A (earthfill) and decomposed content, sometimes very 150 sheepsfoot roller. Spec.
as outer shoulder fill both up and downstream. MMDD). specification. Zone 1A used up and
water added). quartzite. 'rocky'. OMC to 2% dry of OMC.
downstream of central core.

Well-compacted clayey silty sands / sandy Reasonably compacted. Early rockfill


SM (SC/ML/MH) - 4p Cat 834 padfoot and 4p
residual 150 (225 silts/clays with medium to high plasticity fines. Granodiorite, FR to max < 0.7 m, 10 to 25% 4p 10t SDVR, placed dry was dirty (10 to 25% finer than 19 mm
Greenvale 1A (earthfill) Clayey silty sands to 100.3% SMDD groove roller, 0.3% wet of 3 (rockfill) - 1200 to 1800
granodiorite loose) Fines content = 30 to 55%. Used in MW finer than 19 mm (?) leaving quarry). Used in upstream
sandy silts / clays. OMC
downstream shoulder. shoulder.

Rockfill of high compressibility and


GW - sandy gravel, Well-compacted. Formed the bulk of the
relatively high fines content. Zone 4
max size 100 mm, downstream shoulder fill and 30 - 40% of the 4 (& 5) Limestone (poor Zone 4 - 0.6m 4 - 4p 10t SDVR
La Angostura 2 (gravels) river alluvium - 300 4p 10t SDVR Max size 0.5 m. - used adjacent to Zone 2 (up and
<3% fines (60 to 75% upstream shoulder fill. Used up and (rockfill) quality) Zone 5 ? 5 - dumped
downstream) and Zone 5 in the outer
gravel) downstream of the central core.
shoulder regions.
GP/GW to SP/SW - Zone 3A - Max 1.0m, Reasonable (outer rockfill) to good (inner
Well-compacted broad transition zone in the up 1.0 (3A) to 2.0
La Grande 4 select sand and Sandy gravel to 4p 9t SDVR, moisture 10% < 19 mm, Cu ~ 5. 4p 9t SDVR, no water rockfill) compaction. Zone 3 used in the
2B (transition) DR = 85% 450 (loose) and downstream shoulders (outer slopes of 1H 3 (rockfill) granite and gneiss - (3B), 3B toward
(LG4) gravel gravelly sand, < 5% added if required. Zone 3B - max. size of added. outer upstream and downstream
to 1V). Of high moduli . outer slope
fines. 2.0 m. shoulders.

Reasonably to well compacted. No indication if Essentially a broad riprap zone on the


rock pices up to 1 m in
Nurek 3 (gravels) alluvium ? sand to boulder sized - 1000 70t rollers (pneumatic) water added. Formed bulk of fill in 4 (rockfill) - - 3000 to 6000 70t rollers (pneumatic ?) outer shoulders and at the downstream
size.
embankment shoulders. toe.
Appendix F Page F38

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 4 of 10)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Shoulders and/or
Pore Water Pressures in Earthfill Core (Average Instrument Settlement Comment Internal Monitoring
(from end of Steady / Filters
Comments Comment mm/year) Type / No. Comment Equipment
embankment Fluctuating
of mm % of H
construction)
Mid 1975 to Mid 1981.
Vertical strain profile varied from HSGs indicated relatively
First filling started before Annual drawdown of 10 to 15 m. Largest
Excess PWP in core at end of construction. 0.5 - 1.0% in the upper section uniforn settlements
Andong -1.0 to 3.95 years EOC and reached within Fluctuating drawdown of 16 m in 1981/82 (4 - 5 years - IVM 1040 1.25% - HSGs and IHMs.
Slowly dissipated with time over 6 to 8 years. to 2.2% the near base (3.8% in between core and outer
13 m of FSL before post EOC).
contact zone). earth and rockfill zones.
EOC.

Annual drawdown, typically less than 5m. Considered as "normal"


Filling started prior to Relatively small
Larger drawdowns of 8 to 12m in During operation core & foundation respond to behaviour. Vertical strain
end of construction. Fluctuating IVM B in Zone 3 - deformations indicate
Bradbury -0.3 to 5.1 years 1970/72, 1975/77 and 1984/86. Largest reservoir level fluctuation; amplitude decays as 460 IVM, 1 in core 720 1.21% increases with depth from 0 to -
Reached FSL in April (small) / Steady 143mm or 0.38%. relatively high moduli of
drawdown of 26m from 1986 to 1991 (33 move downstream. Minimal lag to reservoir. 1% near crest to 3 to 4% near
1958. compacted rockfill.
to 38 yrs). base.

Initially steady (1969 to 1979).


During construction, PWP only recorded in the Considered as "normal" Zone 1C (IVM A) -
Started in Feb 1966. Fluctuating after 1979, but not an annual
Fluctuating contact zone of Zone 1A (r u ~ 0.3), dissipated in 2 1795 behaviour. Vertical strain 2.60% Similar strain profiles in
Burrendong -0.6 to 6.2 years Reached FSL in Dec drawdown. Large drawdown events (18 - IVM 2.46% -
(slow) years. Otherwise no postive PWP in core during (IVM B) increases with depth from 0 to Zone 1B (IVM C) - Zones 1B and 1C.
1969. to 25 m) in 1979/81, 1981/83, 1993/95
construction. 1% near crest to 5% near base. 2.46%
and 1997/98.

Started in 1987, reached Annual drawdown, typically of 20 to 40m.


FSL in early 1996. Maximum reservoir level varies from year
Canales 0.4 to 10 years Fluctuating - - - - - - - - -
Within 28m of FSL in to year. Rapid rise of 90m over 5 months
early 1990 (3.7 yrs). to FSL in 1996.

Only have records since Annual drawdown, typically 2 to 7 m.


Fluctuating During construction, high pwp in Zone 1 (r u up to
Cardinia unk 1980 (6.5 yrs), filled prior Larger drawdown of 17 m in1982/83 (8.5 1090 - - - - - - -
(small) 0.7). Slow dissipation over more than 26 years.
to this. to 10 yrs).

Annual drawdown, typically up to 20 m,


Started in Mar 1951. but several larger events up to 30 m in
USBR (1997) refer to piezometric levels still rising
Dixon Canyon 2.25 to 4.45 years Reached FSL in May Fluctuating 1954, 1955, 1958 and 1976. 1954 - - - - - - - -
within the dam.
1953. drawdown largest (30 m) and to lowest
elevation.

Possibly stored water


during construction, but IVMs installed in core, but
No annual drawdown. Maintained within Line of equi-potentials indicates uniform head loss
Eucembene -1 to ? not sure to what level. Steady 710 - - - - - - insufficient results to
10 m of FSL. across the core zone.
Unknown when reached analyse.
FSL.

1971 to mid 1973. Delay Steady (after Annual drawdown of 2 to 8 m, reduced to On reservoir fluctuation, ~ 60% of response
Greenvale 0.96 to 3.04 years whilst embankment 1983) / 2 to 5 m after 1983. Largest drawdown recorded in upstream region of the core, but - - - - - - - -
completed. Fluctuating in 1982 (12 - 13 yrs) of 12 m. virtually no response in the central region.

Filling started in May Significant differential


La Angostura 0.4 to >1.1 1974. Within 15 m of - no information. - - IVM - - - - settlement and therefore IVM, HSGs, IHMs.
FSL by Dec 1974. arching across core.

During construction - low PWP in core (ru ~ 0.1 to


La Grande 4 700 (30 to 40% Limited settlement of core
1.3 to 1.75 years. March 1983 to Dec 1983. - - 0.2 max). Almost complete dissipation during IVM - - - - -
(LG4) as snow) during construction.
winter shutdown.

During construction - positive PWP not recorded Internal settlement of Greater settlement of
First filling occurred as
until 15 to 25 m of fill above piezometer. no indication of profile of internal gravel shoulders, upstream shoulder due to
Nurek not known construction proceeded. - no information. Thereafter PWP increased in proportion with - IVM 13700 4.80% -
settlements. Upstream - 11.9 m collapse settlement on
Near FSL at EOC.
increasing embankment height (ru up to 0.93). Downstream 4 to 6.5 m saturation.
Appendix F Page F39

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 5 of 10)
SURFACE MONITORING - POST CONSTRUCTION
Time of Crest Upstream Shoulder Downstream Shoulder
Name Initial General Comments
Readings S LT (% per Comments Comments on Deformation and Observations.
Settlement Displacement Settlement Displacement Settlement Displacement
(years after log cycle of
(mm) (mm) (mm) (mm) (mm) (mm)
EOC) time)

Crest settlement shows an increase in rate (on log scale)


Internal core settlement behaviour is considered normal. Would
after 6 years, which is at odds to the behaviour of the The cause of the acceleration in the crest settlement rate after 6
0 (shoulders) 276 mm or 232 mm or 127 mm or consider as normal except for acceleration in crest settlement after 6
Andong 0.39 - - 10 mm (1.7 yrs) upstream slope. Displacement records are limited, but years is not known. SLT is high in comparison to dams with SM/GM
0.13 (crest) 0.33% (7.5 yrs) 0.34% (7.5 yrs) 0.22% (1.7 yrs) years. Not sure of cause of observed crest settlement behaviour,
indicates downstream displacement of downstream cores.
does not appear to be due to reservoir operation.
shoulder in early years.

Settlement shows large reduction in rate after first filling. Consider as "normal" deformation behaviour for main section of dam. Deformation behaviour of main section of embankment is considered
Acceleration in rate of settlement (all SMPs on main The acceleration in settlement rate appears to have also occurred in "normal". Greater settlements (as a percentage of the embankment
215 mm or 205 mm or 160 mm or
Bradbury 0.23 years 0.06 61 mm (44 yrs) 43 mm (44 yrs) 67 mm (44 yrs) section) between 36 and 41 years, possibly due to the foundation (IVM B). SMPs indicated approx. 50 mm settlement height) on the left abutment between Stations 26 and 28. Settlement
0.33% (44 yrs) 0.32% (44 yrs) 0.37% (44 yrs)
historically large drawdown. Displacement results are over this period whilst IVM B indicates approx. 43 mm. Possibly due of the foundation at IVM B about 30 to 40% of the post construction
erratic but indicate downstream displacement. to a change in effective stress conditions in the foundation. crest settlement at main section.

Consider as "normal" deformation behaviour. Several SMPs show


Settlement is considered normal. Displacement of
0 - slopes and large deformations, but these were indicated as being "loose" or Consider as "normal" deformation case. Note increase in settlement
330 mm or 191 mm or 113 mm or shoulders is upstream for upstream slope and
Burrendong crest sett 0.16 -11 mm (34 yrs) -131 mm (34 yrs) 67 mm (34 yrs) "undermined". Greater settlement (as % of H) of SMPs on the poorly rate (to ~20 mm/yr) of upstream slope on 1997/98 drawdown (the
0.43% (34 yrs) 0.33% (34 yrs) 0.20% (34 yrs) downstream for downstream slope. Displacement of the
0.33 - crest displ compacted shoulders than the core. Settlement rate increase of the most rapid event in dam history).
crest is limited.
upstream shoulder on rapid drawdown in 1997/98.

1009 mm or Longitudinal cracking on crest located a the downstream edge of the


Centre of crest settled some 400 mm during the rapid Case of abnormal deformation behaviour. Vertical scarp developed
0.65% (10.3 clay core, first observed in 1989. Differential settlement across the
filling in 1996 to FSL. Data records to 1992 show large at downstream edge of core of up to 600 mm. Possible development
Canales 0.29 years yrs). Much less - - - - - - crack of up to 600 mm. Indicates collapse settlement of upstream
differential settlement (approx. 500mm) between of shear in core. Rockfill susceptible to collapse settlement on
for downstream rockfill and either large plastic deformation or shear development in
downstream edge of crest and centre of crest. wetting, indicating likely poor placement methods of rockfill.
edge. the wet placed CH clay core.

Gradual reduction in rate of settlement with time (normal


Consider as "normal". Crest displacement is upstream, but this
288 mm or 70 mm or 114 mm or time). Displacements - crest and upstream slope
0.14 on slopes, occurred during the first 2 years (presumably on first filling), and
Cardinia 0.34% 0.19 -65 mm (25 yrs) 0.11% -55 mm (2.5 yrs) 0.18% 134 mm (25 yrs) displaced upstream, downstream slope displaced Consider as "normal" deformation behaviour.
0.43 on crest. thereafter the trend has been virtually constant, with some fluctuation
(25 yrs) (2.5 yrs) (25 yrs) downstream. Crest displacement responds to reservoir
following the fluctuation in reservoir level. No observation of cracking.
fluctuation.

Large magnitude and high long-term rate of settlement. Suspect some influence but limited of foundation on settlement, but
Trend shows a gradual increase in settlement rate for the not possible to quantify. Suspect increase in rate of settlement (on Consider as possibly 'abnormal' deformation behaviour given the
1150 mm or 1320 mm or 590 mm or crest and upper downstream slope. Displacement is log time scale) is possibly due to gradual wetting up of earthfill as large magnitude and on going high rate of settlement. Embankment
Dixon Canyon 0.36 years 1.38 82 mm (39 yrs) -158 mm (47 yrs) 296 mm (47 yrs)
1.80% (39 yrs) 1.97% (47 yrs) 1.14% (47 yrs) upstream for u/str slope, then shift to downstream after 30 indicated by increasing piezometric heads for more than 40 years. raised in 1988 due to large magnitude of crest settlement and to
yrs. Crest and d/stream slope displace d/str with Possible that the dry placed earthfill is susceptible to collapse provide storage for the maximum probable flood.
continuing d/str trend. compression on wetting.

Limited records during early years of operation.


160 mm or 104 mm or Consider as 'normal' deformation behaviour. Displacement show
0.42 vertical Displacement is downstream for crest and downstream Consider as a case of "normal" deformation behaviour. The reservoir
Eucembene 0.14% 0.08 96 mm (40 yrs) - - 0.14% 44 mm (40 yrs) continuing downstream deformation and settlements show slow rate
6.6 horiz. slope (base reading is 6.6 yrs after EOC). Settlements is maintained at a steady level.
(40 yrs) (40 yrs) of settlement.
are small in magnitude.

Consider as normal deformation behaviour. IVM at 25 m high section


0.21 to 0.31 0.0 (when Downstream displacement of the crest and d/str slope on
70 mm or 8 mm or 0.02% shows initial settlement and then a gradual heave post construction.
Greenvale (well before start reservoir 58 mm (28 yrs) - - 54 mm (28 yrs) first filling, continuing at decreasing rate with normal time. Consider as "normal" case
0.13% (28 yrs) (28 yrs) Downstream SMP at main section shows initial heave followed by
of first filling) steady) Settlement pattern similar.
small settlement.

Most of the post construction deformation occurred within Limited deformation data, only 6 months during the period of first
Displacement on first filling indicates crest spreading of 180 to 190
the rockfill, with limited deformation in the gravels (Zone filling. Deformation indicates crest spreading and greater settlement
0.4 (from start of 127 mm or 515 mm or 158 mm or mm. Cracking may have occurred as a result. Larger settlement of
La Angostura - 107 mm (0.6 yrs) -230 mm (0.6 yrs) 121 mm (0.6 yrs) 2). The downstream crest and shoulder displaced of upstream slope (than core and downstream slope). Likely arching
ff) 0.09% (0.6 yrs) 0.38% (0.6 yrs) 0.14% (0.6 yrs) upstream slope possibly associated with collapse compression in the
downstream on first filling, and the upstream crest and between the narrow, wet placed clay core ansd the well compacted
rockfill on wetting.
slope displaced upstream. inner gravel shoulders.

unknown (before
La Grande 4 Limited data. Relatively small downstream displacement Limited data. Very high moduli of core and Zones 2A and 2B likely
first filling, < 1.3 - - 60 mm (2.6 yrs) - - - 63 mm (2.6 yrs) Limited data.
(LG4) on and post first filling. reason for small magnitude displacement on first filling.
yrs)

Reservoir filling occurred as construction proceeded. Records


indicate collapse compression of the upstream sand-boulder fill was
Nurek - - - - - - - - No data. No information
a significant factor in the large magnitude of settlement during
construction.
Appendix F Page F40

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 6 of 10)
GENERAL DETAILS CONSTRUCTION / DESIGN
Construction Timing Embankment Classification Dimensions
Name Owner/ References
Location Geology Foundation Dam Upstream Downstream Comments on Design
Authority Year Time Core Core Height, H Length, Ratio
Zoning Slope Slope
Completed (years) n Width Type (m) L (m) L/H
Class (H to V) (H to V)

Central cut-off to bedrock, up to Very broad earthfill zone with downstream shoulder of
Pueblo Kinney & Bartholomew (1987)
Sedimentary - sandstone and 20m depth. Shoulders founded on compacted weathered rockfill and gravels. Rockfill zone
(right abutment USA, Colorado USBR 1974, Oct (?) 2 4,0,1 c-vb CL 51 1480 29.0 3H to 1V 2.5H to 1V Wright (1987)
shale (with bentonite seams) low plasticity clays and weathered under upstream shoulder and foundation drain under
embankment) USBR (1972)
shales. downstream shoulder.

Broad central cut-off (max. 15m


depth and 45m base width) to
4,0,0 Very broad zoned central earthfill core with thin shoulder Sherard et al (1963)
bedrock with shoulders on 15m
Rector Creek USA, California - - 1946, late - (close to c-vb SC/SM 61 271 4.4 2.5H to 1V 2.25H to 1V zones of gravels (upstream) and rockfill (downstream). Sherard (1973)
depth of pervious soils (sands and
homog 0,0,0) Broad central cut-off of max 15m depth. ICOLD (1974)
gravels ?). Abutments on bedrock
or shallow soil profile.

Pliocene-Pleistocene age non- Over-consolidated clays of low to


marine sediments. Over- high plasticity, clayey sands and Broad central clay core supported by compacted earth and
San Justo USA, California USBR 1986, Jan 0.5 4,2,2 c-tk CL 41 340 8.3 2.5H to 1V 2H to 1V USBR (1995)
consolidated clays, silty sands silty sands of 100's of metres rockfill shoulders. Waste berm at upstream toe.
and clayey sands. depth.

USBR (1974)
Soil - deep alluvials in floodplain 3H to 1V (upper 2.25H to 1V (upper Very broad central core (2H to 1V upstream and 1H to 1V
San Luis - Main Sedimentary, marine - shales, Hickox & Murray (1984)
USA, California USBR area. Bedrock and shallow soil 1967, June 6 4,2,1 c-vb CL 116 5650 48.7 44m), 8H to 1V 44m), 6H to 1V downstream) supported by shoulder zones of compacted
Dam sandstones & conglomerates Morfitt (1984)
profile on hillslopes. below this below this. earth and rockfill.
Von Thun (1988)

USBR (1974)
Very broad central core (2H to 1V upstream and 1H to 1V
Bedrock and shallow soil profile. 3H to 1V (upper Ballard et al (1981)
San Luis - Slide Sedimentary, marine - shales, downstream) supported by shoulder zones of compacted
USA, California USBR Colluvial (slope-wash) and residual 1967, June 6 4,1,1 c-vb CL 30 to 45 - - 44m), 8H to 1V 2.25H to 1V Morfitt (1984)
Area sandstones & conglomerates earth and rockfill. Constructed on upstream sloping hill-
soil profile. below this Von Thun (1982, 1988)
slope.
Stark & Duncan (1987, 1991)

Central core with shoulders of sand & gravel, and rockfill.


Kim (1979)
Soyang Korea - - Bedrock (?) 1972, Dec 5 4,1,0 c-tm GC 123 530 4.3 2H to 1V 2H to 1V Cofferdam forms upstream toe. Filters on upstream and
Sonu (1985)
downstream side of core.

Soil foundation. Low plasticity


sandy clays, clayey sands, clayey Very broad central core with outer shoulders of rock fines
Sedimentary - sandstone &
Spring Canyon USA, Colorado USBR gravels and sandy silts. Up to 15 1948, late (?) 1 (approx) 4,1,0 c-vb CL 68 347 5.1 2.5H to 1V 2.5H to 1V and rockfill. Broad (50 m wide) cut-off to bedrock, aligned USBR (1946, 1997)
shale
m depth in valley and 3 m depth on upstream of dam axis.
abutments. Cut-off to bedrock.

Broad central core with 0.25H to 1V upstream and 1H to 1V


Core on sound rock (except in
downstream slope. Supported by compacted earthfill Hunter et al (1974)
Australia, New upper abutments), Shoulders on 1.5
Tooma SMHEA granitic gniess & biotite granite 1961, March 4,1,2 c-tk SM 67 305 4.6 2.5H to 1V 2.25H to 1V downstream and poorly compacted rockfill upstream. SMEC (1986)
South Wales weathered / decomposed to sound (seasonal)
Filters either side of central core and on downstream GHD (1995b)
rock.
foundation.

Sedimentary (Ordivician) - Removal of alluvium in valley floor


4,1,0
Goulburn Murray mudstones siltstones and to expose HW sedimentary rocks. Very broad central clay core with small rockfill zones SMEC (1996c)
Tullaroop Australia, Victoria 1959, Apr 0.8 (close to c-vb CL 41 427 10.4 2.35H to 1V 2.5H to 1V
Water sandstones, overlain by tertiary Abutments on tertiary alluvials and upstream and downstream HECEC Aust. (1999)
2,1,0)
basalts. basalt.
Appendix F Page F41

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 7 of 10)
MATERIALS
Earthfill Core (Zone 1) Filter or Transition Zones
Name Dry Density Relative
3 Moisture Layer
ASCS (t/m ) or n Density or
Zone Source n Placement Method Content at Comments Zone Source ASCS Class Thickness Placement Method Comment
Class Density Ratio Dry Density
Placement 3 (mm)
(% SMDD) (t/m )

Pueblo Zone 1 makes up approx. 90% of


(Spec. >100% Well compacted - 150 mm (Spec. 0.5% dry to
(right abutment 1 alluvial CL (SC & GC) materials used. Clay to gravel size - - - - - - No filters used.
SMDD) layers, 12p tamping rollers 1.5% dry of OMC)
embankment) materials. Max size 125 mm.

Well compacted clayey sands to silty


Residual / Well compacted - 150 mm sands of low plasticity placed well dry of
Rector Creek 1 weathered SC/SM - layers, 10p sheepsfoot 2 to 4% dry of OMC OMC. Fines content approx. 45%. Zone - - - - - - No filters used.
igneous rock rollers 1 used in central lower core section and
cut-off.

Well compacted filter zones. Dual filter on


Well compacted Zone 1 is central core with 0.5H to 1V up
pleistocene CL (CH,SC, 2A & 2B 2A - SP 2A - 300 downstream side of central core (combined
San Justo 1 & 1A 100% SMDD Zone 1 - 150 mm layers by - and downstream slopes. Zone 1A - D R = 90% vibratory rollers
sediments SM,ML) (filters) 2B - GP 2B - 600 width of 8m) and on downstream foundation.
tamping rollers earthfill zone in downstream shoulder.
Single filter (Zone 2A) used upstream of core.

Well-compacted sandy clays of low to


Likely well compacted. Zone 2 located
San Luis - Main alluvial & Well compacted - 150 mm medium plasticity (66% avg fines sand, gravel &
1 CL (SC) 102% SMDD 1.2% dry of OMC 2 (transition) alluvium - 300 crawler tractors downstream of core over lower half of core
Dam colluvial layers, tamping rollers content) . Very broad central earthfill cobbles
height. Only used in main sections.
zone.

Well-compacted broad central core zone.


San Luis - Slide alluvial & Well compacted - 150 mm Zone 2 not used in embankment section in slide
1 CL (SC) 102% SMDD 1.2% dry of OMC Sandy clay (66% avg fines content) of - - - - - -
Area colluvial layers, tamping rollers area.
low to medium plasticity.

Reasonably to well
3
99.5% SMDD compacted - 350 mm Clayey sandy gravel, no indication of 2.16 t/m 7p 18t pneumatic roller, in Well-compacted filters. Located upstream and
Soyang 1 - GC 3 1.2% wet of OMC 2 (filter) alluvial (?) GW - sandy gravel 400
1.81 t/m layers, 8p 25t tamping plasticity. 20 to 36% fines. D R = 70% moist condition downstream of central earthfill core.
roller

Well compacted. Located downstream of


Mainly sandy clays of low plasticity, GM - 125 mm max, central core over the full dam height and
98.8% SMDD Well compacted - 150 mm Avg 2.9% dry of rockfill (rock 3 10p tamping rollers, wetted
Spring Canyon 1 alluvial CL 3 placed well dry of Standard Proctor 2 (transition) 43% pass 6.75 mm, 2.08 t/m 300 upstream over the upper 43 m. Width
1.78 t/m layers, 12p tamping rollers OMC fines) for compaction
OMC. 13.3% fines increasing with depth (up to 30 m wide at base
of dam).

Well-compacted silty gravelly sand. Problems with achieving grading specification,


completely Well compacted - 150 mm 2 - GW/GP (max
98.7% SMDD Fines (31%) of low plasticity. Particle crushed rock (& 4 passes of crawler type particularly early on due to high fines content.
Tooma 1 weathered SM 3 layers, 12p (minimum) 1.0% wet of OMC 2 (filter) size 40 mm, avg 8% - 150 to 305
1.72 t/m breakdown during compaction shifted sand blend) tractor Filters located on up and downstream side of
granite sheepsfoot roller fines)
fines content and OMC. core and on downstream foundation.

3 Limited information. Located between clay core


Tullaroop 1 - CL 1.76 t/m Known to be rolled - limited information. 2 (filter) - gravels - - -
and rockfill toe, up and downstream .
Appendix F Page F42

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 8 of 10)
MATERIALS
Either Earthfill or Rockfill Zone Either Earthfill or Rockfill
Name
Dry Layer Layer
Grading Grading Dry Density
Zone Source Density Thickness Placement Method Comments Zone Source 3 Thickness Placement Method Comments
Description 3 Description (t/m )
(t/m ) (mm) (mm)

3 - sandstone 3 - spread, no formal Poorly compacted Zone 3 used under


Pueblo sand, gravel & Well compacted. Zone 2 used in outer
4p 18t crawler tractors, 3 and 4 4 - limestone, shale 3 - coarse 3 - 900mm compaction upstream shoulder.
(right abutment 2 (gravels) alluvial cobbles, max size 300 Spec D R >75% 300 downstream shoulder (inner slope 2H to 1V, -
watered (rockfill) and sandstone 4-? 4 - 300 mm 4 - 12p tamping rollers, Well compacted Zone 4 used the in
embankment) mm outer slope 2.5H to 1V)
(weathered ?) placed close to OMC inner downstream shoulder zone

Limited information. Zone 3 (gravels)


Same source to Zone 1 but coarser grading.
Residual / used in outer upstream shoulder, and
10p sheepsfoot rollers, 2 Zone 2 used as broad fill zone in upper portion 3 (gravels),
Rector Creek 2 (earthfill) weathered igneous SC/SM ? - 150 - - - - - Zone 4 (rockfill) used in outer
to 4% dry of OMC of embankment and up and downstream of 4 (rockfill)
rock downstream shoulder. Both relatively
Zone 1.
thin shoulder zones.

4 - decomposed Reasonably to well compacted. Zone 4


4 4-?
Zone 3 used in upper and outer upstream granite 4-? used in upstream shoulder (upstream of
San Justo 3 (rockfill) limestone - - 1000 vibratory rollers 1A 1A - 100% 300 pneumatic tyred rollers
shoulder. Suspect well-compacted. 1A - pleistocene 1A - CL central core). In the downstream
(both earthfill) SMDD
sediments shoulder, Zone 4 encapsulates Zone 1A.

No formal compaction of rockfill zones. Zones


Reasonably to well compacted.
4 and 5 used in the mid to upper region of the
San Luis - Main 4 - 300 mm 3 (earth and alluvial, colluvial & tamping rollers, placed Miscellaneous fill zone used in lower
4 & 5 (rockfill) basalt clean ? - crawler tractor upstream shoulder. Zone 4 used downstream variable 99% SMDD 300
Dam 5 - 900 mm rockfill) weathered rock (mix) 1.4% dry of OMC upstream shoulder and in outer
of central core and along downstream
downstream shoulder.
foundation as a filter zone.

No formal compaction of rockfill zones. Zones


Reasonably to well compacted.
4 and 5 used in the mid to upper region of the
San Luis - Slide 4 - 300 mm 3 (earth and alluvial, colluvial & tamping rollers, placed Miscellaneous fill zone used in lower
4 & 5 (rockfill) basalt clean ? - crawler tractor upstream shoulder. Zone 4 used downstream variable 99% SMDD 300
Area 5 - 900 mm rockfill) weathered rock (mix) 1.4% dry of OMC upstream shoulder and in outer
of central core and along downstream
downstream shoulder.
foundation as a filter zone.

GW - sandy gravel 3 Poor compaction. Limited information.


2.16 t/m 5p 10t SDVR. In moist Well-compacted. Inner earthfill zone upstream
Soyang 3 (gravels) alluvial (?) with 1 to 4% fines 600 4 (rockfill) - - - 1000 Trafficked by dump truck Placed in outer downstream and
D R = 70% condition. and downstream of the central core.
(80% gravel) upstream shoulders.

Sedimentary - not No indication of rock type or strength. Zone 3


no formal compaction, no
Spring Canyon 3 (rockfill) specific on which gravel to boulder size - 900 used as outer fill zone in up and downstream - - - - - - -
water added.
type shoulders.

3000
Poorly compaction. Zone 3 used in upstream Reasonable compaction of highly
upstream weathered granite 92.6% to >98% 12p sheepsfoot rollers,
Tooma 3 (rockfill) granite ? - - spread and sluiced shoulder. Rockfill also used in lower and very 1A (earthfill) variable 150 variable material. Zone 1A used in
900 to 3000 (?) - CW to HW SMDD average 1% dry of OMC
outer slope of downstream shoulder. downstream shoulder.
downstream

Limited information. Located at upstream and


Tullaroop 3 (rockfill) basalt - - - - - - - - - - -
downstream toe regions.
Appendix F Page F43

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 9 of 10)
HYDROGEOLOGY MONITORING - DURING CONSTRUCTION
First Filling Reservoir Operation Core
Name Rainfall Other Types of
Timing Shoulders and/or
Pore Water Pressures in Earthfill Core (Average Instrument Settlement Comment Internal Monitoring
(from end of Steady / Filters
Comments Comment mm/year) Type / No. Comment Equipment
embankment Fluctuating
of mm % of H
construction)
Started Jan 1974,
Pueblo Foundation settlement
reached FSL in Sept Annual drawdown, typically less than 10 Base plates under upstream
(right abutment -0.75 to 11 years Fluctuating - - - - - - - under upstream shoulder
1985 (close to FSL in to 15 m shoulder.
embankment) of 370 to 460mm.
1983 & 1984).

Seasonal drawdown of 15 to 20 m.
Approx. 0 to 2.5 Suspect negligible pore pressures in core during
Rector Creek Jan 1947 to May 1949 Fluctuating Limited data, only first 3.5 years of - - - - - - - -
years construction due to placement well dry of OMC.
operation.

During construction- No positive PWPs indicating


likely placement of the core on the dry side of
OMC. Base plates indicate fndn
Started February 1986,
San Justo 0.1 to 2.1 years Fluctuating annual drawdown of up to 20m During operation - piezometric levels continued to - - - - - - settled 76 mm during 1 base plate, 1 inclinometer
reached FSL Feb 1988.
rise (1986-1998) and have not reached steady construction.
conditions. Core shows some but limited
response to reservoir fluctuation.

No records from during construction. Large foundation


Annual seasonal drawdown typically up During operation - response (timing and settlements for
San Luis - Main Started Sept 1967, 2120 (B) 2.44% (B) Increasing strain with increasing base plates, IHMs, IVMs
0.25 to 2 years Fluctuating to 40 m, larger events - 1981 of 65 m, amplitude) to reservoir fluctuation dependent on - IVM, 2 (B & D) - embankment section
Dam reached FSL June 1969 3060 (D) 3.14% (D) depth below crest. (upstream shoulder)
1994 of 50+m & 1997 of 50 m. the piezometer location. Negligible response in over alluvial floodplain
centre of Zone 1. deposits.

No records from during construction.


During operation - response (timing and
Annual seasonal drawdown typically up
San Luis - Slide Started Sept 1967, amplitude) to reservoir fluctuation dependent on
0.25 to 2 years Fluctuating to 40 m, larger events - 1981 of 65 m, - - - - - - - -
Area reached FSL June 1969 piezometer location. Negligible response in centre
1994 of 50+m & 1997 of 50 m.
of Zone 1. Elevated pore pressures in Zone 1 and
soil foundation on large and rapid drawdown.

Sept 1972 to Oct 1975. Likely arching in narrow


Significantly lower
First filling started shortly High PWP in core at end of construction (ru up to Vertical strain profile varied from clayey gravel core.
magnitude of
Soyang -0.3 to 2.85 years before end of Fluctuating Annual drawdown of 15 to 25m. 0.7). Dissipated within 1 to 2 years. Limited - IVM 3800 3.10% 1.0% in the upper section to 4.5 Relatively Large plastic IHMs.
settlement in the sandy
construction. 85 m response to reservoir level post construction. to 6% the near base. deformations in the mid
gravels.
below FSL at EOC. section of core.

Annual drawdown, typically up to 20 m,


Started in Apr 1951. but several larger events up to 30 m in
USBR (1997) refer to piezometric levels still rising
Spring Canyon 2.4 to 4.6 years Reached FSL in May Fluctuating 1954, 1955, 1958 and 1976. 1954 - - - - - - - -
within the dam.
1953. drawdown largest (30 m) and to lowest
elevation.

Vertical strain profile varied from


0.5 to 1.0% in the upper section IVM in core located
Approximate estimation Subject to rapid drawdowns of up to 30 During operation - gradual loss of head and
Tooma 0 to 0.2 years Fluctuating - 1 (core) 1527 2.24% to 3.5 to 4.5% the near base - slightly upstream of dam -
of time of first filling. m. Daily fluctuation is up to 3 m/day. reduced respose to reservoir level across core.
(6.3% at contact). Considered axis.
normal.

No records of reservoir
operation in first 9 years. Annual drawdown of 5 to 8 m. Larger
Limited response of piezometers within the core to Data from IVMs not available for
Tullaroop 0 to ? Suspect FF by late 1960 Fluctuating drawdowns of 10 to 11.5 m in 1983, 490 - - - - - -
changes in reservoir level. construction period.
to mid 1961 (1.5 to 2 1995, and 1998.
yrs).
Appendix F Page F44

Table F1.2: Zoned earth and rockfill embankments case study information (Sheet 10 of 10)
SURFACE MONITORING - POST CONSTRUCTION
Time of Crest Upstream Shoulder Downstream Shoulder
Name Initial General Comments
Readings S LT (% per Comments Comments on Deformation and Observations.
Settlement Displacement Settlement Displacement Settlement Displacement
(years after log cycle of
(mm) (mm) (mm) (mm) (mm) (mm)
EOC) time)

Settlements show decreasing rate with time (normal


Pueblo 0.88 years 455 mm or 340 mm or Consider as 'normal' deformation. Displacement readings are erratic.
58 mm (22.5 yrs), 295 mm or -35 mm (22.5 yrs), 160 mm (22.5 yrs), time). Displacement records are erratic, general trends Central concrete buttress flanked by two earthfill embankments.
(right abutment (reservoir 36m 0.80% (22.5 0.61 0.99% (22.5 There is some consistency with fluctuating reservoir level, but limited
readings erratic 0.76% (22.5 yrs) erratic readings erratic readings are upstream slope deforms upstream, crest and Deformation post construction is considered 'normal'.
embankment) below FSL) yrs) yrs) consistency between the crest and downstream shoulder.
downstream shoulder deform downstream.

Cracking first observed in Feb 1947, 18mm wide diagonal crack


across crest persistent for depths > 7m. New cracks in Dec 1949 Consider as a case of 'abnormal' deformation in terms of settlement
Large magnitude crest settlements and high settlement
(end of 16m drawdown) at each end of embankment, also diagonal and displacement trends. Collapse compression of the dry placed
855 mm or rate post FF. Displacements initially upstream (approx.
Rector Creek 0.1 to 0.2 years 1.65 140mm (10 yrs) - - - - across crest, 12mm width. Cracking and displacement behaviour earthfill is a likely significant factor in the deformation behaviour
1.74% (10 yrs) 240mm max.) for most of first filling and then sharp
(ICOLD 1974) due to gradual wetting up of dry placed earthfill. Large observed. Likely brittleness of the dry placed earthfill was probably a
change to downstream at 2 to 2.2 years.
settlements likely a result of collapse compression on wetting of the significant factor contributing to the cracking.
dry placed earthfill. Possible fndn influence also.

Relatively small post construction deformations. Base plate shows


Deformations of small magnitude. Settlement shows
fndn heave of 45 mm since EOC, most of which occurred during FF,
'normal' trend. Displacements < 10mm on first filling.
which has probably reduced the amount of settlement measured by
48 mm or 40 mm or 35 mm or Long-term, crest shows very slow d/stream trend as does Consider embankment deformation behaviour as 'normal'. Slide in
San Justo 0.25 yrs 0.125 2 mm (10 yrs) -22 mm (8.5 yrs) 11 mm (10 yrs) the SMPs. Following the earthquake in Oct 1989, SMPs on shoulders
0.11% (10 yrs) 0.15% (8.5 yrs) 0.09% (10 yrs) d/stream slope, upstream shows u/str trend. left abutment away from dam.
settled some 15 mm (crest negligible) and SMPs displaced upstream
Deformations affected by Loma Prieta earthquake in
some 10 to 20 mm. Some crest cracking, but not sure if this is
1989.
related to the earthquake.

SMPs taken at Stn 110 where fndn is bedrock or a Differential displacements on first filling indicate crest spreading of Consider deformation behaviour at main section as 'normal'. Crest
shallow soil depth. SMPs show relatively large settlement 225 mm at Stn 110. Crest spreading observed for SMPs along a spreading and cracking was observed along large portions of the
415 mm or 415 mm or 195 mm or
San Luis - Main 165 mm (Stn 100, -55 mm (Stn 100, 185 mm (Stn 100, on FF, thereafter behaviour is 'normal' . Upstream slope large portion of the embankment length. Longitudinal cracking on the embankment, and was not confined to the area of the upstream slide
0 to 0.08 years 0.38% (Stn 0.26 0.43% (Stn 110, 0.34% (Stn 110,
Dam 30 yrs) 30 yrs) 30 yrs) displaces u/str on first filling then gradual d/str trend. crest observed (as a result of crest spreading) along large portions of in 1981. Suspect wetting of the dry placed earthfill as well as the
110, 30 yrs) 30 yrs) 30 yrs)
Crest and d/str slope displace downstream on FF with the crest. Larger deformations (both settlement and displacements) foundation and hill-slope topography were influences on the crest
gradual downstream trend, more so for the d/str slope. were observed for the embankment section over the soft alluvium. cracking.

Significant difference in settlement between the upstream


Settlement behaviour on large drawdowns (prior to slide) shows
and downstream sides of the crest. Displacement is
acceleration in settlement of SMPs 13m upstream. Plotting
158 mm or upstream for SMPs on crest and upstream slope, and
130 mm or settlement as % of dam height highlights the deformation at SMP 136 Slide in upstream slope occurred during the large drawdown in 1981.
San Luis - Slide -65 mm (Stn 138, 0.53% (Stn 136, -183 (Stn 136, 14 shows crest spreading. SMPs on upstream slope (13m
1 to 0.08 years 0.41% (Stn 0.35 - - as abnormal (in the slide vicinity). This may be an indicator of the Consider the deformation behaviour of the SMPs on the upstream
Area 30 yrs) 14 yrs, prior to yrs, prior to slide) upstream) show acceleration in settlement during the
138, 30 yrs) impending failure, although the natural slope topography is likely to slope in the vicinity of the slide as 'abnormal'.
slide) large drawdowns at years 8 and 10. This is exaserbated
have an influence. The SMP at 13m upstream at Stn 136 was
at Stn 136 when the settlement is plotted as % of
located just above the backscarp of the slide.
embankment height.

Consider as normal deformation. IVM indicates post construction


360 mm or Relatively small magnitude settlements post construction settlements in core of up to 360 mm after 5 years, with most of the
62 mm or
Soyang 0 0.29% (5 yrs) 0.095 - - - - as measured by SMPs. IVM indicates greater settlement settlement (70%) within the lower third of the embankment. Possibly Consider as 'normal' deformation behaviour.
0.08% (5 yrs)
from IVM. of core. related to consolidation as the high construction pore water pressures
dissipated.

Large magnitude and long-term rate of settlement,


particularly the crest and upstream shoulder. Settlement Suspect some but limited influence of the foundation on settlement,
Consider as possibly 'abnormal' deformation behaviour given
rate (in log time) trend shows an increase with time for but cannot quantify. The increase in settlement rate (on log time
0.25 years 900 mm or 955 mm or 420 mm or magnitude and long-term rate of settlement. Embankment raised in
Spring Canyon 1.31 -18 mm (39 yrs) -45 mm (47 yrs) 160 mm (47 yrs) crest the and upper d/stream slope. Displacement is u/str scale) is possibly due to gradual wetting up of earthfill as indicated by
(approx.) 1.50% (39 yrs) 1.54% (47 yrs) 1.13% (47 yrs) 1988 due to large magnitude of crest settlement and to provide
for crest and u/str slope, then shift to downstream after 20 increasing piezometric heads for more than 40 years. The dry placed
storage for the maximum probable flood.
to 30 yrs. D/stream shoulder displaces d/str with earthfill may be susceptible to collapse compression.
continuing d/str trend.

Settlement trends look normal, apart from acceleration of


Consider as 'normal' deformation behaviour. Unlikely that the
upstream slope from 13 to 16 years. Displacement of Consider as "normal" deformation behaviour. Zoned earth and
acceleration in settlement of the upstream slope is drawdown related
121 mm or 142 mm or 47 mm or crest and upstream shoulder is upstream on first filling rockfill embankment with upstream shoulder of poorly compacted
Tooma 0.10 years 0.12 -32 mm (38 yrs) -41 mm (38 yrs) 8 mm (38 yrs) as the embankment had been subject to several rapid 30 m
0.20% (38 yrs) 0.21% (38 yrs) 0.09% (38 yrs) with continued upstream deformation post first filling. and sluiced rockfill. Subject to rapid fluctuations in reservoir level (up
drawdowns previously. Not sure of cause, but effect it not evident in
Downstream slope displaces downstream on FF then to 3 m/day).
internal core settlement profile.
steady to upstream trend.

Consider as 'normal' deformation behaviour. Internal core settlement


Displacement of upstream shoulder was downstream for profile (from 1979 to 1998) shows no indication of plastic of shear
138 mm or 88 mm or 171 mm or Consider as "normal" deformation behaviour. Embankment could be
Tullaroop 1.32 years 0.23 86 mm (40 yrs) -68 mm (40 yrs) 84 mm (40 yrs) the first 20 years, then deformed upstream. The rest of zones. Change in displacement of upstream shoulder may indicate
0.33% (40 yrs) 0.29% (40 yrs) 0.51% (40 yrs) classified as earthfill with rock toe (2,1,0).
embankment deformed downstream. cracking that has developed under the cyclic operation of the
reservoir.
Appendix F Page F45

Table F1.3: Zoned earthfill embankments case study information


GENERAL DETAILS CONSTRUCTION / DESIGN
Construction Timing Embankment Classification Dimensions
Name Owner/ References
Location Geology Foundation Dam Upstream Downstream Comments on Design
Authority Year Time Core Core Height, H Length, L Ratio
Zoning Slope Slope
Completed (years) n Width Type (m) (m) L/H
Class (H to V) (H to V)

Boyle (1965)
Core on bedrock. Shoulders on
Broad central gravelly core supported by Jones (1965)
Meridian Sedimentary - argillite & recent alluvial gravels in the river
Benmore New Zealand 1964, May 4 3,2,1 c-tk GM/GC 110 823 7.5 2.5H to 1V 2.2H to 1V compacted gravelly shoulders. Some Tait (1965)
Energy greywacke section (20 to 22 m depth and
rockfill used in the downstream shoulder. Opus (1998a)
medium dense to dense, D R > 65%)
WCS (1996a)

Granodiorite, meta- Very broad central clay/silt earthfill core


Goulburn 9 (shutdown State Rivers (1983)
Cairn Curran Australia, Victoria sediments in upper left Bedrock. 1956, Apr 3,0,0 c-vb CL/ML 44 656 14.9 2.75H to 1V 2.7H to 1V with sands and gravels in the outer
Murray Water for ~ 3yrs) Hydro Technology (1995)
abutment shoulders.

Broad central core with relatively flat


Igneous - magnesite and Jones (1955)
Cobb New Zealand Trans Alta Rock. 1955, Mar - Apr 2.3 3,0,1 c-tk GM/GP 35 213 6.1 3H to 1V 2.75H to 1V shoulders of sandy gravels (upstream) and
serpentine. Glaciated valley. Oborn (1985)
talus downstream. No embankment filters.

Very broad central earthfill zone with up


and downstream zones of compacted Rao (1957)
Hirakud India - - Foundation stripped to bedrock 1956 4 (seasonal) 3,0,1 c-vb SC-GC 59 24300 412 3H to 1V 2.25H to 1V
earthfill. Rockfill at downstream toe and an Rao & Wadhwa (1958)
upstream facing of riprap.

Very broad central clay core with outer


Soil foundation. Low plasticity sandy
shoulders of compacted sands and
clays, clayey sands, clayey gravels
Sedimentary - sandstone & gravels, and weathered rockfill. Transition
Horsetooth USA, Colorado USBR and sandy silts. Up to 10 to 20m 1948, Dec 0.6 3,0,0 c-vb CL 48 567 11.8 3H to 1V 2.5H to 1V USBR (1946, 1997)
shale of rock fines on upstream side of core.
depth in valley, thin mantle on
Broad (30 m width) cut-off to bedrock
abutments. Cut-off to bedrock.
upstream of dam axis.

Public Power Katzias & Stamatopoulos (1975)


Core on bedrock. Shoulders on Narrow central core with shoulders of well-
Kastraki Greece Corp. of - 1969, May 1.75 3,1,0 c-tn CL 95 450 4.7 2.25H to 1V 1.75H to 1V Coumoulos & Koryalos (1979)
alluvium in gully (approx 8 m depth) compacted gravelly sands.
Greece Pagano et al (1998)

Central & downstream zone of silty sand


earthfill with gravels in upstream shoulder. SMHEA (1998b)
Australia, New Alluvial deposits overlying Core and shoulders (at main section)
Khancoban SMHEA 1965 1.3 3,0,2 c-vb SM 18 1067 59.3 3.25H to 1V 2.25H to 1V Transition zone on upstream of core and Howard et al (1974)
South Wales metamorphic bedrock on gravels.
filters on foundation. Cut-off slightly P.J. Burgess & Assoc. (1991)
upstream of axis.

Public Power
Sedimentary - limestone, Narrow central core with shoulders of well- Coumoulos (1979)
Kremasta Greece Corp. of Karstic foundation. 1965, Sept - 3,1,0 c-tn CL 165 - - 2.5H to 1V 2H to 1V
karstic foundation compacted sands and gravels Breznik (1979)
Greece

Abutments on granite bedrock. In Broad core (1.1H to 1V upstream and near


Southern
narrow river section part of core on vertical downstream slopes) supported by ENR (1960)
Mammoth Pool USA, California California Granite 1959, Oct 1 3,2,2 c-tk SM 113 250 2.2 3.5H to 1V 2H to 1V
bedrock (cut-off), shoulders on river shoulders of compacted earthfill. Rockfill Wilson (1973)
Edison Co.
sands and gravels (up to 30m depth) toe.

Over-consolidated glacial deposits,


Broad central core (1.75H to 1V upstream
Glacial activity - morraine some alluvial deposits. Cut-off to
slope and 0.5H to 1V downstream)
deposits overlying Tertiary bedrock where glacial deposits
Meeks Cabin USA, Wyoming USBR 1970, Aug 4 (seasonal) 3,1,1 c-tk CL 57.5 1005 17.5 3H to 1V 2.25H to 1V supported by outer gravel shoulder zones. USBR (1998)
aged clay shales with pre- relatively shallow (maximum section
Broad (30m wide) cut-off trench to bedrock
sheared bentonitic seams. and right abutment). Pre-sheared
located upstream of dam axis.
bentonitic seams in shale.
Appendix F Page F46

Table F1.3: Zoned earthfill embankments case study information (Sheet 2 of 10)
MATERIALS
Earthfill Core Filters / Transition Zone
Name Density Ratio, Relative
ASCS or Moisture n Density, or
Zone Source n Placement Method Comments Zone Source ASCS Class Placement Method Comment
Class Dry Density Content Dry Density
3 3
(t/m )

You might also like